Sei sulla pagina 1di 256

Mapping

Degree Theory

Enrique Outerelo
Jesus M. Ruiz

Graduate Studies
in Mathematics
Volume 108

American Mathematical Society


Real Sociedad Matematica Espanola
Mapping Degree
Theory

Enrique Outerelo
Jesus M. Ruiz

Graduate Studies
in Mathematics
Volume 108

American Mathematical Society


Providence, Rhode Island

Real Sociedad Matematica Espanola


Madrid, Spain
Editorial Board of Graduate Studies in Mathematics
David Cox, Chair
Steven G. Krantz Rafe Mazzeo Martin Scharlemann

Editorial Committee of the Real Sociedad Matematica Espanola


Guillermo P. Curbera, Director
Luis AlIas Alberto Elduque
Emilio Carrizosa Pablo Pedregal
Bernardo Cascales Rosa Maria Mir6-Roig
Javier Duoandikoetxea Juan Soler

2000 Mathematics Subject Classification. Primary OlA55, 01A60, 47Hll, 55M25, 57R35,
58A12, 58J20.

For additional information and updates on this book, visit


www.ams.org/bookpages/gsm-108

Library of Congress Cataloging-in-Publication Data


Outerelo, Enrique, 1939-
Mapping degree theory / Enrique Outerelo and JesUs M. Ruiz.
p. cm. - (Graduate studies in mathematics j v. 108)
Includes bibliographical references and index.
ISBN 978-0-8218-4915-6 (alk. paper)
1. Topological degree. 2. Mappings (Mathematics). I. Ruiz, JesUs M. II. Title.
QA612.098 2009
515'.7248-dc22
2009026383

Copying and reprinting. Individual readers of this publication, and nonprofit libraries
acting for them, are permitted to make fair use of the material, such as to copy a chapter for use
in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Requests for such
permission should be addressed to the Acquisitions Department, American Mathematical Society,
201 Charles Street, Providence, Rhode Island 02904-2294 USA. Requests can also be made by
e-mail to reprint-permission«lams. ~rg.
© 2009 by the authors.
Printed in the United States of America.
@) The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at http://www.ams.org/
10 9 8 7 6 5 4 3 2 1 14 13 12 11 10 09
To Jesus Maria Ruiz Amestoy
Contents

Preface ix

I. History 1
1. Prehistory 1
2. Inception and formation 14
3. Accomplishment . . . . . 28
4. Renaissance and reformation 35
5. Axiomatization ... 38
6. FUrther developments 42

II. Manifolds 49

1. Differentiable mappings . 49
2. Differentiable manifolds . 53
3. Regular values . . . . . 62
4. Thbular neighborhoods 67
5. Approximation and homotopy 75
6. Diffeotopies . 80
7. Orientation . 84

III. The Brouwer-Kronecker degree 95


1. The degree of a smooth mapping . 95
2. The de Rham definition . . . . . . 103
3. The degree of a continuous mapping . 110

-vii
viii Contents

4. The degree of a differentiable mapping 114


5. The Hopf invariant ....... . 119
6. The Jordan Separation Theorem. 124
7. The Brouwer Theorems . . . . . . 133

IV. Degree theory in Euclidean spaces 137


1. The degree of a smooth mapping. . . 137
2. The degree of a continuous mapping . 145
3. The degree of a differentiable mapping 153
4. Winding number. . . . . . . 156
5. The Borsuk-Ulam Theorem. 160
6. The Multiplication Formula 171
7. The Jordan Separation Theorem. 176

V. The Hopf Theorems 183


1. Mappings into spheres. . . . . . . . . . . . . . . 183
2. The Hopf Theorem: Brouwer-Kronecker degree. 191
3. The Hopf Theorem: Euclidean degree 196
4. The Hopf fibration. . . . . . . . . . 200
5. Singularities of tangent vector fields 206
6. Gradient vector fields . . . . . . . . 211
7. The Poincare-Hopf Index Theorem. 218

Names of mathematicians cited 225

Historical references 227

Bibliography 233

Symbols 235

Index 239
Preface
This book springs from lectures on degree theory given by the authors
during many years at the Departamento de Geometria y Topologia at the
Universidad Complutense de Madrid, and its definitive form corresponds
to a three-month course given at the Dipartimento di Matematica at the
Universita di Pisa during the spring of 2006. Today mapping degree is a
somewhat classical topic that appeals to geometers and topologists for its
beauty and ample range of relevant applications. Our purpose here is to
present both the history and the mathematics.
The notion of degree was discovered by the great mathematicians of
the decades around 1900: Cauchy, Poincare, Hadamard, Brouwer, Hopf,
etc. It was then brought to maturity in the 1930s by Hopf and also by
Leray and Schauder. The theory was fully burnished between 1950 and
1970. This process is described in Chapter I. As a complement, at the
end of the book there is included an index of names of the mathematicians
who played their part in the development of mapping degree theory, many
of whom stand tallest in the history of mathematics. After the first his-
torical chapter, Chapters II, III, IV, and V are devoted to a more formal
proposition-proof discourse to define and study the concept of degree and
its applications. Chapter II gives a quick presentation of manifolds, with
special emphasis on aspects relevant to degree theory, namely regular val-
ues of differentiable mappings, tubular neighborhoods, approximation, and
orientation. Although this chapter is primarily intended to provide a review
for the reader, it includes some not so standard details, for instance con-
cerning tubular neighborhoods. The main topic, degree theory, is presented
in Chapters III and IV. In a simplified manner we can distinguish two ap-
proaches to the theory: the Brouwer-Kronecker degree and the Euclidean
degree. The first is developed in Chapter III by differential means, with a
quick diversion into the de Rham computation in cohomological terms. We
cannot help this diversion: cohomology is too appealing to skip. Among
other applications, we obtain in this chapter a differential version of the
Jordan Separation Theorem. Then, we construct the Euclidean degree in
- ix
x Preface

Chapter IV. This is mainly analytic and astonishingly simple, especially


in view of its extraordinary power. We hope this partisan claim will be
acknowledged readily, once we obtain quite freely two very deep theorems:
the Invariance of Domain Theorem and the Jordan Separation Theorem,
the latter in its utmost topological generality. Finally, Chapter V is de-
voted to some of those special results in mathematics that justify a theory
by their depth and perfection: the Ropf and the Poincare-Ropf Theorems,
with their accompaniment of consequences and comments. We state and
prove these theorems, which gives us the perfect occasion to take a glance
at tangent vector fields.
We have included an assorted collection of some 180 problems and exer-
cises distributed among the sections of Chapters II to V, none for Chapter
I due to its nature. Those problems and exercises, of various difficulty, fall
into three different classes: (i) suitable examples that help to seize the ideas
behind the theory, (ii) complements to that theory, such as variations for
different settings, additional applications, or unexpected connections with
different topics, and (iii) guides for the reader to produce complete proofs
of the classical results presented in Chapter I, once the proper machinery
is developed.
We have tried to make internal cross-references clearer by adding the
Roman chapter number to the reference, either the current chapter number
or that of a different chapter. For example, III.6.4 refers to Proposition 6.4
in Chapter III; similarly, the reference IV.2 means Section 2 in Chapter IV.
We have also added the page number of the reference in most cases.
One essential goal of ours must be noted here: we attempt the simplest
possible presentation at the lowest technical cost. This means we restrict
ourselves to elementary methods, whatever meaning is accepted for ele-
mentary. More explicitly, we only assume the reader is acquainted with
basic ideas of differential topology, such as can be found in any text on the
calculus on manifolds.
We only hope that this book succeeds in presenting degree theory as
it deserves to be presented: we view the theory as a genuine masterpiece,
joining brilliant invention with deep understanding, all in the most accom-
plished attire of clarity. We have tried to share that view of ours with the
reader.

Los Molinos and Majadahonda E. Outerelo and J.M. Ruiz


June 2009
Chapter I

History
In the body of mathematics, the notion of degree stands as a beautiful achievement
of topology and one of the main contributions of the twentieth century, which has
been called the century of topology. In Chapter I we try to outline how the ideas
that led to this fundamental notion of degree were sparked and came to light. It is
only natural that such a task is biased by our personal opinions and preferences.
Thus, it is likely that a specialist in, say, partial differential equations would present
the tale in a somewhat different way. All in all, a choice must be made and ours
is this:

§l. Prehistory: Gauss, Cauchy, Liouville, Sturm, Kronecker, Poincare, Picard,


Bohl (1799-1910).
§2. Inception and formation: Hadamard, Brouwer (1910-1912).
§3. Accomplishment: Hopf, Leray, Schauder (1925-1934).
§4. Renaissance and reformation: Nagumo, de Rham, Heinz (1950-1970).
§5. Axiomatization: Fiihrer, Amann, Weiss (1970-1972).
§6. Further developments: Equivariant theory, infinite dimensions.

The presentation of these topics is mainly discursive and descriptive, rigorous


proofs being deferred to Chapters II through V where there will be complete
arguments for all the most classical results presented here.

1. Prehistory
Rougly speaking, degree theory can be defined as the study of those tech-
niques that give information on the existence of solutions of an equation
of the form y = f(x), where x and y dwell in suitable spaces and f is a
map from one to the other. The theory also gives clues for the number
of solutions and their nature. An important particular case is that of an
equation x = f(x), where f is a map from a domain D of a linear space
into D itself: this is the so-called Fixed Point Problem.
- 1
2 1. History

By its very nature, it is clear that the origins of degree theory should
be traced back to the first attempts to solve algebraic equations such as

Zn + alzn-l + ... + an = 0,
where the coefficients ai are complex numbers, an i= O. That such an
equation always has some solution is the Fundamental Theorem of Algebra.
This result was most beloved by KARL-FRIEDRICH GAUSS (1777-1855),
who found at least four different proofs, in 1799, 1815, 1816, and 1849.
It is precisely in the first and fourth proofs where we find what can be
properly considered the first ideas of topological degree. By some properties
of algebraic curves (which were formalized only in 1933 by ALEXANDER
OSTROWSKI (1893-1986)), Gauss was able to prove that inside a circle of
big enough radius, the algebraic curve corresponding to the real part of
the polynomial shares some point with the algebraic curve corresponding
to the imaginary part. In this way the following two lines of research were
born:
Problem I. Find the common solutions of the equations

f(X' y) = 0,
{
F(x,y) = 0
inside a given closed planar domain, on whose border the two functions
f(x, y) and F(x, y) do not vanish simultaneously.
Problem II. Find the number of real roots of a polynomial in one variable
with real coefficients, in a given closed interval [a, bJ of the real line.

***
The first contributions to Problem I are due to AUGUSTIN LOUIS
CAUCHY (1789-1857). In a memoir presented before the Academy of Turin,
on November 17,1831, and in the paper [Cauchy 1837a], Cauchy introduces
a new calculus that, in its own words, can be used to solve equations of any
degree.
Some parts of Cauchy's arguments are not completely precise, and the
way these parts were made rigorous by JACQUES CHARLES FRANQOIS
STURM (1803-1855) and JOSEPH LIOUVILLE (1809-1882) is quite relevant
in the history of the analytic definition of the topological degree of a con-
tinuous mapping.
Let us describe this. The definition of the index of a function given by
Cauchy in [Cauchy 1837aJ is the following:
1. Prehistory 3

Let x be a real variable and f (x) a function that becomes infinite at


x = a. If the variable x increases through a, the function will either change
from negative to positive or change from positive to negative or not change
sign at all. We will say that the index of f at a is -1 in the first case,
+1 in the second, and 0 in the third. We define the integral index of f
between two given limits x = Xo and x = Xl, denoted by J:~ (J), as the sum
of the indices of f corresponding to the values of x in the interval [xo, Xl]
at which f becomes infinite. If f is a function in two variables, we define
the integral index of f between the limits xo, Xl; Yo, YI to be the number

~~Jff~ (J) = ~ [J:~ (J(., YI)) - J:~ (J(., YO)) - Jff~ (J(XI' .)) + Jff~ (J(xo, .))] .
In his 1831 memoir, Cauchy obtained the index of a function by integral
techniques and residues and proved the following result:
Theorem. Let r be a closed curve that is the contour of an area S, and
let Z(z) = X(x, y) + iY(x, y) be an entire function. Then

_1 r Z'(z) d =! JS=s" (X/Y)


27l'i } r Z(z) z 2 S=S'

is the number of zeros of Z(z) in S; here s stands for the arc length along
r, and s" - s' is the length of r.
Cauchy generalized this result in a memoir published June 16, 1833, in
Thrin. The generalization follows:
Theorem. Let F(x, y) and f(x, y) be two functions of the variables x, y,
continuous between the limits X = Xo, x = XI, Y = Yo, Y = YI. We denote
by q;(x, y), ¢(x, y) the derivatives of the functions with respect to x, and by
1f/(x, y), 'IjJ(x, y) their derivatives with respect to y. Finally, let N be the
number of the different systems of values x, y, between the above limits,
verifying simultaneously the equations F(x, y) = 0, f(x, y) = O. Then

where
f(x,y) ( )
Ll(x,y) = F(x,y) q;(x,y)'IjJ(x,y) -1f/(x,y)¢(x,y)

= f(x, y) (8F(X, y) 8f(x, y) _ 8F(x, y) 8f(x, y))


F(x, y) 8x 8y 8y f:)x

An elementary "proof" of this theorem appears in [Cauchy 1837b].


However, Liouville and Sturm in [Liouville-Sturm 1837] give three examples
4 1. History

showing that the second theorem above above can fail. The first example
is
F (x, y) = X2 + y2 - 1,
{
f(x,y) = y.
In this example
2xy
L1(x,y) = x 2 + y2 - l'

and drawing around the origin a rectangle containing the circle x 2 + y2 = 1,


one sees that
~~Jff~(L1) = 0,
because L1 never becomes infinity on the sides of the rectangle. However,
the system
X2 + y2 - 1 = 0,
{
y=o
has the two solutions (1,0), (-1,0) inside the rectangle. Liouville and
Sturm conclude their note with the following remark:
There is a theorem that can replace Cauchy's. Let us consider a closed
contour r on which F(x, y) and f(x, y) do not vanish simultaneously, and
let us also assume that inside this contour the function

w = q;(x, y)1jJ(x, y) -l}/(x, y)¢(x, y)


8F(x, y) 8f(x, y) 8F(x, y) 8f(x, y)
8x 8y 8y 8x

does not vanish at the values (x, y) at which f(x, y) and F(x, y) vanish.
In this situation, among the solutions (x, y) of the equations F(x, y) = 0,
f(x, y) = 0, inside r, some correspond to positive values of wand others
to negative values of w. We denote by J-Ll the number of solutions of the
first kind, and by J-L2 the number of solutions of the second kind. With this
notation we have
1
"2 8 = J-Ll - J-L2,
where 8 stands for how many more times the function ~t~',~ changes from
positive to negative than from negative to positive, at those points in the
contour r at which that function becomes infinite, when the contour is
traced in the positive direction.
We see that the function w is the Jacobian of the mapping (F, f) (Li-
ouville and Sturm always consider entire functions). Consequently, we find
1. Prehistory 5

displayed here for the first time the importance of the sign of the functional
determinant
aF aF
w=
ax ay
af af
ax ay
when dealing with the computation of the number of solutions of the system

F(X'Y) = 0,
{
f(x,y) =0
in a planar region.
Today, in the hypotheses of the Liouville-Sturm Theorem, the number
J.LI - J.L2 is called the topological degree of the mapping (F, f) at the origin,
and this is the starting point for the analytic definition of degree. But this
will not take full shape until 1951.

***
In the later paper [Cauchy 1855], Cauchy states the Argument Principle,
which is another way to compute the indices he has defined earlier. These
results, translated into more modern terminology, read as follows.
Winding number (or index) of a planar curve around a point. Let
r c C be a closed oriented curve with a CI parametrization:
z(t) = x(t) + iy(t) + a, 0 ~ t ~ 1, z(O) = z(l), aEC\ r.
Then,

w(r, a) = _1
211"i
r~ = ~ Jor x(t)y'(t)
Jr z - a
l - x'(t)y(t) dt
x 2(t) + y2(t)
211"

is an integer.

This integer is called the winding number (or index) of r around a.


Geometrically, the winding number tells us how many times the curve
wraps around the point. In case r is only continuous, the winding number
is defined through a CI approximation r l of r, because w(n, a) remains
constant for rl close enough to r.
The following example illustrates this notion:
6 1. History

__....- ....-.-=...--- --'--- ..... :~~'-~~.~. ;e


0

I
/ B\ \
I \
I \
\

To proceed one step further, Cauchy considers a simply connected do-


main Gee (that is, G has no holes), a holomorphic function f : G -t C,
( = f(z), and a C1 closed curve reG, on which f has no zeros. Then:

Argument Principle. The following formula holds:

1 /,
w(J(r), O) = -2·
7r~
d(
--;-
f(r) ..,
1
= -2·
7r~
1
r
f' (z)
-f() dz
z
=L
k
w(r, ak)Qk,

where the ak's are the zeros of f in the domain D bounded by r and the
Qk'S are their respective multiplicities.
Suppose next that r has no self-intersection and that it has the positive
(counterclockwise) orientation. Then D is a connected domain (this is the
Jordan Separation Theorem, which we will discuss later), and w(r, a) = +1
for all a ED, so that the last formula becomes

w(J(r), O) = LQk,
k

that is:
Theorem. The total number of zeros (counted with multiplicities) that f
has in D is the winding number of the curve f(r) around the origin.
1. Prehistory 7

In general, the winding number can be negative, but we can still say
that f has at least Iw(f(r), 0)1 zeros in the domain bounded by r.

***
Let us now turn to Problem II. The first full solution is due to Sturm.
In 1829 and 1835 he gave an algorithm to find the exact number of dis-
tinct real roots of a polynomial. The theorem was later generalized by
CARL GUSTAV JACOB JACOBI (1804-1851), CHARLES HERMITE (1822-
1901), and JAMES JOSEPH SYLVESTER (1814-1897).
Exploring the topological content of Sylvester's article [Sylvester 1853],
LEOPOLD KRONECKER (1823-1891) introduces in his papers [Kronecker
1869a] and [Kronecker 1869b] a method that extends Sturm's. Indeed, at
the end of his work Kronecker writes:
In my research developed in this article, I started from a theorem by
Sturm. A generalization of that result was found by Hermite some time ago,
but I have been able to extend the continued fraction algorithm developed by
Sylvester to further widen Sturm's theorem.
Let us describe Kronecker's contribution. He starts with the following
definition:
Regular function systems. A regular function system consists of n +1
real functions Fo, FI, ... , Fn in n real variables Xl, ... , X n , such that

(a) Fo, F I , ... , Fn are continuous and have no common zeros. They admit
partial derivatives with respect to all n variables, and those deriva-
tives take finite values.

(b) The functions Fo, F I , ... ,Fn take positive and negative values. More-
over, each function takes positive (resp., negative) values infinitely
often.

(c) The domains {Fi < a}, i = 0, ... , n, represent n-dimensional varieties
that only contain finite values of the variables Xl, ... , X n .

(d) No functional determinant

I~~I k;ii=O,I, ... ,n' k=O,l, ... ,n,


j=l, ... ,n

vanishes at any zero of the system Fk 1= 0, Fo = FI = ... = Fn = 0.


8 1. History

(e) The common zero set of any chosen n-l functions among Fo, FI, ... ,
Fn is a CI curve.

Then Kronecker looks at the orientations of the CI curve involved in


this definition (condition (e) above). He considers this part basic in his
research on systems of functions in several variables:
Orientation Principle. Kronecker chooses for every pair (k, i), k < i, an
orientation of the CI curve (recall (e) above)

F(k, i) = {x ERn: Fi(X) =0 for i =f k, £}.

This orientation is denoted by Ikil; he then puts likl = -Ikil.


Next, he defines:

(a) A point e E F(k,i) n {Fk = O} is called an incoming (eingang) point


of F(k, i) (in {x E R n : Fk(X) . Fe(x) < O}) if the following condition
holds true: walking the curve F(k, i) as oriented by Ikil, we leave the
set {x E R n : Fk(X) . Fe(x) > O} at the point e and enter {x E Rn :
Fk(X) . Fe(x) < O}.
The set of all these incoming points e is denoted by E(k, i).

(b) A point a E F(k,i) n {Fk = O} is called an outgoing (ausgang) point


of F(k,i) (off {x ERn: Fk(X) . Fe(x) < O}) if the following condition
holds true: walking the curve F(k,i) as oriented by Ikil, we leave the
set {x E R n : Fk(X) . F£(x) < O} at the point a and enter {x E R n :
Fk(X) . F£(x) > O}.
The set of all these outgoing points a is denoted by A(k, i).

After the preceding preparation, Kronecker shows that the number

#E(k, i) - #A(k, i)

is even and does not depend on the indices k, i, and then he defines:
Kronecker characteristic. The characteristic of the regular function
system Fo, FI , ... ,Fn is the integer

X(Fo, FI, .. . ,Fn) = H#E(k, i) - #A(k, i)).

It is convenient to stress that in the course of his proof of this fact


1. Prehistory 9

Kronecker obtains the following very modern description of his invariant:

(Fo, . .. , Fn)(x) =0
Fj(x) <0

= (-1)j .
(Fo, .. . ,Fn)(x) = 0
Fj(x) > 0

(here J stands for the Jacobian determinant) for any j = 0, ... , n.


Once this invariant is defined, Kronecker shows how it detects solutions
of the given regular system:
Kronecker Existence Theorem. Let Fo, FI, ... , Fn be a regular function
system. If x(Fo, F I , ... , Fn) -=1= 0, then for every i = 0,1, ... , n, the system

Fo(x) = 0, ... , Fi-I(X) = 0, Fi(X) < 0, Fi+1(X) = 0, ... , Fn(x) = °


has some solution x E ]Rn.

This result and the above remark extend to a system of n functions


with n unknowns what Liouville and Sturm had done thirty years earlier,
as has already been mentioned.
We illustrate all of this with a simple example:

Fa =0
10 I. History

Here, we depict a regular system consisting of three functions Fo, F l , F2 :


]R2--+ R Each set

is a Cl curve, and we have the incoming points E(0,2) = {el, e2, e3} and
the outgoing points A(0,2) = {all. We get
1 1
X(Fo,FI,F2) = 2(#E(0,2) - #A(0,2)) = 2(3 -1) = 1 f= 0,

and we see immediately in the picture that each of the three systems

(0) Fo < 0, FI = 0, F2 = 0,
(1) Fo = 0, FI < 0, F2 = 0,
(2) Fo = 0, FI = 0, F2 < °
indeed has solutions.

In 1877, influenced by some discussions with his friend KARL THEODOR


WILHELM WEIERSTRASS (1815-1897) on the matter of complex analysis
and potential theory, Kronecker gave a representation of his characteristic
by means of an integral, today known as the Kronecker integral. Indeed,
he shows this:
Kronecker Integral Theorem. Let Fo, F I , ... , Fn be a regular system.
Then Jor every j = 0,1, ... , n,
1 ( ~
X(Fo,FI, ... ,Fn ) = - vol(§n-l) Js i[>n du
J

where
Fo(w) FI(W) --
FiJ:!!) Fn(w)
QEb. !Hl ~ 8Fn
~= 8Xl 8Xl 8Xl 8Xl

~
8X n -l
8H
8X n - l
-- 8Fj
(3x n -l
8Fn
8Xn-l

Note that in case n = 2, and supposing that FI and F2 are the com-
ponents of a holomorphic function J(z) on Fo :S 0, the above integral for
°
j = becomes
_1 /, f'(z) dz.
27ri r J(z)
1. Prebistory 11

Thus, the Kronecker Existence Theorem is in fact a generalization of Cau-


chy's results.
There are antecedents to the Kronecker integral in dimension> 2. Even
Kronecker remarks in [Kronecker 1869a] that a special case of his integral
can be found in a previous paper by Gauss on potential theory [Gauss 1813].
In this case n = 3,

and the Kronecker integral gives the flow through the surface Po = 0 of the
electric field
x - xO
V(z) = Ilx _ xOl13
created by the unit charge placed at xO = (x~, xg, xg). Gauss establishes in
this paper that the flow is equal to 471" or 0, according to whether the point
xO is interior or exterior to the surface. In modern terms, the flow through
the surface Fo = 0 is expressed by

Lo=o (V, 1/)dS,


where 1/ is the outward normal vector field on the surface and dS is the
area element.

***
After Kronecker's pioneering work on the characteristic and the integral
of a regular system, there were a number of papers on the subject, with
important applications to geometric and topological questions. Following
the chronological development, me must mention first the contributions by
JULES HENRI POINCARE (1854-1912), who used the Kronecker integral
in the qualitative theory of autonomous (independent of time) ordinary
differential equations. Already in 1881, Poincare uses the index of a planar
curve to study these differential equations, but in 1883 he uses Kronecker's
theorem in a note in Comptes Rendus [Poincare 1883]. There he writes:
Mr. Kronecker has presented to the Berlin Academy, in 1869, a mem-
oir on functions of several variables, includin9 an important theorem from
which the following result follows easily:
Let 6,6, ... ,~n be n continuous functions in the n variables Xl, X2, ... ,
x n , the variable Xi restricted to range among the limits -ai and +ai. Let
us suppose that for Xi = ai the function ~i is always positive, and for
12 1. History

Xi = -ai the function eiis always negative. Then I say that there is a sys-
e
tem of values for the X at which all the vanish. This result can be applied
to the three-body problem to prove it has infinitely many special solutions
with important properties that we are to present.
This is an Intermediate Value Theorem in arbitrary dimension. Three
years later Poincare published the paper [Poincare 1886] in which he stud-
ies the curves defined by differential equations of the second order. There
he looks for the singular points of those equations and discusses their dis-
tribution using the Kronecker integral. This contains the argument for the
invariance of the characteristic under continuous deformations, which is
used in the proof of the result stated above. Kronecker had himself con-
sidered this invariance, and actually he could prove it in some particular
cases.

***
Another important question implicit in Kronecker's work is the deter-
mination of the exact number of solutions of a system of equations. But it
is CHARLES EMILE PICARD (1856-1941) who in 1891 published a note in
Comptes Rendus [Picard 1891] with the precise formulation of the problem
whose resolution was to be his main contribution to degree theory. Picard
states:
Let us consider n equations fi(x}, X2," ., xn) = 0, i = 1,2, ... , n,
where we suppose the Ii's represent continuous functions in n real vari-
ables Xl, X2, ... ,Xn defining a point in some domain D. The question of
finding the number of common roots of these equations in that domain has
held the attention of geometers for a long time. A formula has been given
in this sense by Mr. Kronecker, in his famous investigations of the char-
acteristic of function systems. Unfortunately, the Kronecker integral, a
multiple integral of order n - 1 on the surface of the domain D, does not
give the number of roots we are looking for. The functional determinant of
the system plays a fundamental role in this theory, and one only obtains
the difference of the numbers of roots at which the determinant is positive
and the number of roots at which the determinant is negative.
I will show here, in a few lines, that it is possible to represent the exact
number of roots by a suitable integral.
Picard's method consists of attaching new equations to get a new system
that has the same roots and whose functional determinant is always positive
and applying Kronecker's theory afterwards. Picard presented his results in
1. Prehistory 13

full in classes at La Sorbonne, Paris, and finally published them in a two-


volume treatise on analysis [Picard 1891/1905]. After this, Kronecker's
theory became a classic.
In 1904 PIERS BOHL (1865-1921) published a paper [BohI1904] where
he proves the following result:
Theorem. Let (G) be the domain defined by -ai ::; Xi ::; ai (ai > 0, i =
1,2, ... , n). There do not exist continuous real-valued functions FI, ... , Fn ,
without common zeros in (G), such that Fi = Xi (i = 1,2, ... , n) on the
boundary of (G).
For the proof, Bohl uses the Kronecker integral and Stokes' Theorem
(proving along the way the Kronecker Existence Theorem). From the above
proposition, and by continuous deformation techniques, Bohl deduces the
following two statements:
Proposition. (a) Let (G) be the domain defined by -ai ::; Xi ::; ai (ai >
0, i = 1,2, ... , n), and let FI"'" Fn be continuous real-valued functions
such that Fi = Xi (i = 1,2, ... , n) on the boundary of (G). Then, for
every point ('Y1,T'2, ... ,T'n) in (G) there is another (XI,X2, ... ,Xn ) such
that Fi(XI, X2,· .. , xn) = T'i (i = 1,2, ... , n).
(b) Let (G) be the domain defined by -ai ::; Xi ::; ai (ai > 0, i =
1,2, ... , n), and let iI, ... , fn be continuous real-valued functions that do
not vanish simultaneously in (G). Then there are a constant N < 0 and a
point (UI, ... , un) in the boundary of (G) such that Ii (UI, U2, ... , un) = N Ui
(i=I,2, ... ,n).
It is possible that these results were already in the Ph.D. thesis de-
fended by Bohl in 1900 at Dorpat University. Note here that (b) implies
what is today the well-known fact that the boundary of (G) is not a re-
tract of (G), which in turn implies a fixed point result. Thus we can say
that Bohl was really close to discovering the Brouwer Fixed Point Theorem.

On the other hand, in 1895 Poincare published his famous memoir Anal-
ysis Situs [Poincare 1895b], where, starting from the attempts by GEORG
FRIEDRICH BERNHARD RIEMANN (1826-1866) and ENRICO BETTI (1823-
1892), he starts the development of what will later be called combinatorial
topology and simplicial homology. Poincare refined these ideas in two com-
plements to his Analysis Situs, [Poincare 1899, Poincare 1900], where we
find the first notion of (abstract) polyhedron, which will later be called tri-
angulation of a compact manifold V of dimension n. This was a finite set
T of disjoint cells of different dimensions::; n such that:
14 I. History

(1) Every cell of dimension d, or d-cell, is the image in V of an open


ball B C ~d via a diffeomorphism from an open neighborhood of the
closure B into a submanifold W of V of dimension d.

(2) The boundary in W of such a d-cell is a union of cells of dimensions


::;d-l.

(3) V is a union of cells of T.

We will come back to this, when ten years later the notion of manifold
comes to full development.

2. Inception and formation


In 1910 JACQUES SALOMON HADAMARD (1865-1963) published a remark-
able paper [Hadamard 1910], which is in fact an appendix to the book
[Tannery 1910]. This paper marks the transition from the origins of degree
theory to the construction of a more elaborate and systematic theory.
Hadamard says:
The proof given by Mr. Ames for the Jordan Theorem in the plane is
based on the index (or variation of the argument) of a planar curve with
respect to a point (the Cauchy concept). The generalization of this concept
to higher dimensions is provided by the Kronecker index. This concept is
now classic, mainly after the Traite d'Analyse by Mr. Picard (vol. I, p.
123; vol. II, p. 193) of 1891 and 1893, and after several contemporary
papers that give new applications of that index.
Later Hadamard mentions Poincare, Bohl, and LUITZEN ECBERTUS
JAN BROUWER (1881-1966).
Let us now describe things more explicitly. The result is the following
well-known statement:
Jordan Separation Theorem. Let r be a Jordan curve in ~2, that is,
a curve homeomorphic to a circle. Then ~2 \ r has exactly two connected
components whose common boundary is r.
The argument that Hadamard refers to in his comment runs as follows:

(1) First, by some geometric and topological means, one sees that ~2 \ r
has at most two connected components and only one unbounded.
2. Inception and formation 15

This reduces the result to its most essential part, that r disconnects
the plane.

(2) Second, one computes the index with respect to r. This gives:

(a) Points at which the index vanishes. These belong to the un-
bounded component of ]R2 \ r, which is called the exterior of
r.
(b) Points at which the index is ±1 . These are in a second (bounded)
component, the interior of r.

The figure below illustrates the second (and most relevant) part of this
proof.

I
I
,I
I
I
I
I
I
I

0;~:)~:---------------- ---
CAMILLE JORDAN (1838-1922) stated and tried to prove this theorem
in [Jordan 1893]. As mentioned above, the essential content of the result
is that a Jordan curve divides the plane, which jointly with the fact that a
simple arc does not divide the plane is the oldest proposition of set topology
in Euclidean spaces. The first complete proof of the Jordan Theorem was
given by OSWALD VEBLEN (1880-1960) in [Veblen 1905].

In [Schonfiies 1902]' ARTHUR MORITZ SCHONFLIES (1853-1928) gives


the following additional information concerning this theorem:
16 1. History

Schonflies Theorem. Let r be a Jordan curve in jR2 and C I its interior.


Then, for all x E C I and all a E r, there is a simple arc from x to a whose
points other than a are all in C I .
We say that a is accessible from C I . This finally led to the following
characterization of Jordan curves:
If a compact set in jR2 has two complementary domains, from which
every point of the set is accessible, then the set is a Jordan curve.
Here we see the typical topological beast which is not a Jordan curve:

***
Let us now take a closer look at Hadamard's paper. In the first para-
graph, Hadamard analyzes the proof above of the Jordan Separation The-
orem, focusing on the part involving the order (= index) of a point with
respect to the curve. He thus presents clearly what he wants to general-
ize. To do that, he defines, in the second paragraph, what he means by a
surface in the Euclidean space. His notions are based on ideas introduced
by Poincare, at the beginning of the century, to treat polyhedra. We will
not go into detail here, but in a sketchy way, we can think of an (n - 1)-
surface in jRn as a subspace that decomposes into pieces homeomorphic
to the standard (n - 1)-simplex, which glue in a suitable way along their
faces. Hadamard concludes this paragraph by stating Green's Theorem,
which reduces a volume (triple) integral to a surface (double) one.
In the third paragraph, Hadamard defines the order of a point with
respect to a closed oriented surface in jRn. Let us suppose we are given a
hypersurface S, with coordinates (Xl, ... ,Xn ) E jRn with respect to param-
eters (UI,'" ,Un-I), such that the Xi'S have continuous partial derivatives
with respect to the uj's (later, this condition is weakened to the mere exis-
tence of partial derivatives); suppose also that the origin is not a point of
S. Then, Hadamard shows this key fact:
2. Inception and formation 17

Hadamard Integral Theorem. The following integral is an integer:

w
1 r ~
= - vol(§n-l) } s ,n dUI ... dUn-1

where

~= and , = Jx~ + ... + x~ .

Hadamard calls this integer w, the order of the origin with respect to
S. The order with respect to an arbitrary point a = (al,"" an) ¢ S is
obtained by translation, that is, by replacing Xi with Xi - ai everywhere in
the formula above.

The proof of this theorem proceeds by induction, using Green's Theo-


rem, the starting step being the order (= index) of a point with respect to
a planar curve. It is a long proof: seven pages in the paper! Along the way,
Hadamard obtains several important byproducts. For instance:
Proposition. Let S be a hypersurface, and consider a point a ¢ S. If
there exists a half-line starting at a and not meeting S, then the order of S
with respect to a is zero.
This of course corresponds to the unbounded component of IRn \ S (if
we advance the Jordan Separation Theorem in dimension n). But among
other things, Hadamard explains how the order can be computed:
Proposition. Let S be a hypersurface, which is a closed polyhedron. Let
w stand for the order with respect to S of a point a ¢ S. Then

where NI, N2 are computed as follows. Pick a half-line f starting at a and


meeting S only at points lying in (n - 1) -simplices (not in faces). Then
NI + N2 is the number of points in f n S, and such a point is counted in
NI (resp., N2) if the n vertices of the simplex that contains it, ordered in
the corresponding orientation and preceded by the point a, determine an
n-simplex oriented according to the orientation of the coordinate system.
18 1. History

This is much simpler than it reads. Consider the following polygonal


curve in the plane:

We have three points PI, P2, P3 in f n S, and to decide which is in NI and


which in N2, we construct the three simplices

Comparing with the standard orientation of the coordinate system


a2

~Dal
we immediately see that Pt E N2, P2 E NI, and P3 E N2. Hence,
w(S, a) = 1 - 2 = -1.

Furthermore, Hadamard analyzes the variation of w as a moves and


concludes:
Proposition. (a) The order is constant if the point varies without crossing
the surface.
(b) The order is ±1 if S is a convex polyhedron and the point is an
interior point.
2. Inception and formation 19

In the fourth paragraph, Hadamard defines the Kronecker index of a


function system for functions defined on a closed hypersurface:
Definition. Let S be a hypersurface of the Euclidean space ~n, and let
II, ... , f n be a system of n continuous functions defined on S, which do not
vanish simultaneously at any point of S. The index of the system is the
order of the origin with respect to the hypersurface (II, ... , fn)(S) generated
by the functions.
Note that this definition has meaning by the non-vanishing condition,
which guarantees that the origin does not belong to the hypersurface
(II,···, fn)(S).
The main property of this index is given in the following theorem:
Boundary Theorem. Let S be the boundary of a domain VeRn, and let
II, ... , fn be defined and continuous on the whole domain V. If II,···, fn
do not vanish simultaneously at any point of V, then the index w of the
system II, ... , fn on S is zero.
This can be trivially rewritten as a solutions existence statement:
Corollary. Let S be the boundary of a domain VeRn, and let II,· .. ,fn
be defined and continuous on S, such that they do not vanish simultane-
ously at any point of S. If the index of II, ... , fn on S is not zero, then
the system II = 0, ... , fn = 0 has some solution in V.

On the other hand, the computation of the index of a function system


on a surface depends on the following important result:
Poincare-Bohl Theorem. Let us consider, on the same closed hypersur-
face S C R n , two function systems II, ... , f nand gl, ... , 9n, both satisi.ying
the non-vanishing condition on S. Then:
(i) If the systems have different indices, then there is at least one point
x in S such that

(ii) If the indices of the two systems are not in the ratio (-1)n, then there
is at least one point x in S such that

Using this, Hadamard obtains:


20 1. History

Schonflies Theorem. Consider the closed disc

and its boundary §l. Let f, g : D2 --+ IR be two continuous functions such
that the mapping (f, g) : D2 --+ 1R2 is injective, and let C be the closed curve
defined by the parametrization (flsl, gISl). Then, for every point (X, Y)
interior to C, there is a point (x,y) E D2 such that (f(x,y),g(x,y)) =
(X,Y).
Note how the Jordan Separation Theorem for plane curves is used here.
Moreover, Hadamard states the Jordan Separation Theorem in IRn without
proof and then deduces the Sch6nfiies theorem in IR n :
Schonflies Theorem in IRn. Let V be a domain with boundary a surface 8
in one piece (this hypothesis is not essential), and let h, . .. ,fn be contin-
uous functions from V into IR such that the mapping f = (h,···, fn) :
V -+ IRn is injective. Consider the surface 8' parametrized by fls =
(his, ... , fnls). Then for every point (Xl, ... , Xn) interior to 8', there
is a point (Xl, ... , Xn) in V such that f(xI,"" Xn) = (XI, ... , Xn).
As a consequence, Hadamard gets:
Proposition. If f : V --+ V' is a homeomorphism between two domains,
the interior points of V map to interior points of V', and the boundary
points of V map to boundary points of V' .
This is a particular case of the Invariance of Domain Theorem. Con-
cerning the Jordan Separation Theorem in IR n , it is very likely, as we will
see later, that Hadamard had known it from some private communication
with Brouwer.

Another very important result that Hadamard proves follows:


Brouwer Fixed Point Theorem. Every continuous mapping from the
closed balllf c IRn into itself has at least one fixed point.

Proof. The argument runs as follows. Let f : If --+ If be a continuous


mapping. We can suppose f(x) =1= X for all x E 8lf = §n-l (otherwise we
are done). Then, by the Poincare-Bohl Theorem Idsn-l - flsn-l and Idsn-l
have the same index: were it

Xl - h(x) = ... = Xn - fn(x) = J.L < 0,


Xl Xn
2. Inception and formation 21

then f(x) = (1 - J..L)x tJ. if. Consequently, the index of Idsn-l - flsn-l is
1, and by the basic properties of indices, there is an Xo E if such that
Xo - f(xo) = 0, that is, f(xo) = Xo. 0

Another remarkable statement (without proof) by Hadamard is the fa-


mous Poincare-Hop! Index Theorem, relating the Poincare index (charac-
teristic) of a domain to the zeros of a vector field. In particular, Hadamard
proves:
Vanishing of tangent fields. Let V be a domain that admits a tangent
space at every point, depending continuously on the point. If the index u of
V is not zero, then it is not possible to attach continuously a tangent line
to V at every point (that is, the tangent line must be undefined somewhere).

This is the case, for instance, for the sphere in ]R.3. The fact that
Hadamard formulates this theorem seems to result from a misunderstanding
with Brouwer, who had proved it for spheres.
To conclude, Hadamard states (again without proof) the following the-
orem due to Brouwer:
Proposition. Every continuous mapping from a sphere into itself that
preserves orientations has at least one fixed point.

The result is false if the mapping reverses orientation (think of the


antipodal isometry)

***
One should mention here a footnote (p. 476 of [Hadamard 1910]) where
Hadamard explains that the method for the proof of the previous result
was communicated by Brouwer. Thus, it becomes more and more evident
that there was a quite fluent exchange of ideas between Hadamard and
Brouwer. What follows is a letter from Brouwer to Hadamard that spares
any further comment on the matter. We translate from [Brouwer 1976]:

Paris, January 4, 1910


6 Rue de l'Abbe de l'Epee
Dear Sir,
I can now communicate some extensions of the fixed point theorem for bijective con-
tinuous transformations of the sphere. They are reduced to arbitrary continuous trans-
formations of the sphere. To such a transformation one can attach a finite number n as
its degree. From a degree n transformation one can construct by continuous variations
any other degree n transformation, but no more than those. In particular, one can always
construct in this way a degree n rational transformation of the complex sphere.
22 1. HistOIY

To determine this degree, we use homogeneous coordinates (in the double sense),
write x, y, z for the initial sphere, and ~,TJ, ( for the image, split the sphere into a finite
number of regions, and consider firstly the transformations defined by relations

€: TJ: (= lI(x,y,z): h(x,y,z): h(x,y,z),

where II, h, fa are polynomials, which on the other hand can vary for the different regions
of the sphere. We call this transformation a polynomial transformation. We choose an
orientation on the sphere: then every point P in the image, in general position, will occur
rp times with the positive orientation, and Sp times with the negative orientation. In this
situation, one can prove that rp - Sp is a constant: it. is the degree of the polynomial
transformation.

Let us come back to an arbitrary continuous transformation. It can be approxi-


mated by a sequence of polynomial transformations: one proves then that the latter have
all the same degree: it is again the degree of the limit transformation.

The degree is always a finite integer, either positive or negative. The degree of a
bijective transformation is +1, if the orientation is preserved, and -1, if the orientation is
reversed.

Now the generalized fixed point theorem becomes what follows: Every continuous
transformation of the sphere, whose degree is not -1, has at least one fixed point.

Moreover, I have extended this theorem to m-dimensional spheres. It reads then


in the following way: Every continuous transformation of the m-dimensional sphere has at
least one fixed point, except a) when m is odd and the degree n is +1, b) when m is even
and the degree n is -1.

In particular, if the transformation is bijective [in margin, strenger formuleeren]'


there is at least one fixed point a) if m is odd and the orientation is reversed, b) if m is
even and the orientation is preserved.

For the volume of an m-dimensional sphere [sic] in the space of dimension m + 1 (if
we include there the sphere itself) I was able just recently to establish a still more general
theorem, namely: Every (possibly not bijective) continuous transformation of the volume
of the m-dimensional sphere has at least one fixed point.

Concerning general continuous tangent distributions on the sphere, two articles of


mine will appear soon, where I study certain questions that refer to the Dirichlet principle
and to the decomposition of a field into a "quellenfrei" part and a "wirbelfrei" part. To that
end, I determine first the most general form that tangent curves (characteristic curves after
Poincare) can have. As the main result in the first article, one should take the property
that a characteristic curve that does not tend to a singular point must be a spiral, whose
two limit cycles are also characteristic. The property that there is at least one singular
point is necessary; it is not in the end an accessory corollary, on which I have insisted
because it was the first easy to formulate result and because there seemed to be some close
relation between this theorem and that of the fixed point in the sphere, a relation that
has been clarified only through your correspondence. In the second article I have included
your beautiful, direct, and more complete proof of the existence of at least one singular
point.

My address will be in Paris till January 15. Maybe there will be the occasion for
us to meet?

Yours sincerely,
L.E.J. Brouwer

***
2. Inception and formation 23

Now we turn to Brouwer's fundamental work. We can agree that the


greatest contribution due to Brouwer is the definition of the degree of a con-
tinuous mapping of n-manifolds solely by geometric and topological means.
The manifolds he considers come from the ideas concerning polyhedra that,
as was already mentioned, Poincare introduced at the end of the nineteenth
century and the beginning of the twentieth century. These ideas were the
germ of what we know today as combinatorial topology. It is remarkable
that Brouwer never mentions Poincare's writtings on the topic.
The notion of a manifold used by Brouwer coincides with what nowadays
we call a pseudomanifold. A manifold of dimension n, or n-manifold, is built
up from simplices of dimension d, or d-simplices, of smaller dimensions d.
Skipping strict formalisms, we can put it as follows:
Definition. A subspace M of a Euclidean space is an n-manifold when it is
homeomorphic to the polyhedron IKI associated to a simplicial complex K
of dimension n. That is, IKI is a subspace of a Euclidean space consisting
of a union of disjoint simplices of dimension :S n (points, open segments,
open triangles, open tetrahedrons, etc.) such that;

(i) Every simplex of IKI is a face of some n-simplex of IKI.


(ii) Every (n - I)-simplex of IKI is a face of at most two n-simplices of
IKI·
(iii) For every pair ofn-simplices 8 and 8' oflKI, there is a finite sequence
8 = 8 1,82 , ... , 8m = 8' of n-simplices of IKI such that 8 i and 8i+1
have a common (n - I)-face for i = 1, ... , m - 1.


The boundary of IKI, which will be denoted by IKI, is the unwn of all

(n-l)-simplices that are faces of exactly one n-simplex. Hence, if IKI = 0,
every (n - I)-simplex is a face of exactly two n-simplices of IKI.

The following examples are very simple illustrations of these notions:


(a) A 2-manifold with boundary.

M=D 2 JKJ
24 I. History

(b) A I-manifold without boundary.


IKI=0
(c) A space which is not a polyhedron.

Once the notion of a manifold is fixed in this form, Brouwer constructs


the topological degree of a continuous mapping of manifolds in [Brouwer
1912aj, article dated in Amsterdam, July 1910. In a footnote in the first
page we read:
While this paper was in print, the note by J. Hadamard, Sur quelques
applications de l'indice de Kronecker, has appeared in the second volume of
J. Tannery's Introduction it la tMorie des fonctions d'une variable. In that
note some aspects of the theory we present here are anticipatedly developed.
This once again confirms the mutual influence between the two mathe-
maticians.
Let now M and N be two n-manifolds, which we assume to be con-
nected, compact, boundaryless , and oriented. For the definition of degree,
Brouwer proceeds in two steps:
Step I: The simplicial case. Consider a simplicial continuous mapping 9 :
M -+ N. This means that 9 is continuous, and there are triangulations
r.p: IKI == M and 'IjJ : ILl == N of M and N such that the localization

9 = 'IjJ 09 0 r.p-1 : IKI -+ ILl


restricts to an affine map on each simplex of IKI. For such a 9, Brouwer
shows that there is a dense connected open set [} in ILl such that if y E [}
is not in the union S of the images by 9 of the (n - I)-simplices of IKI,
then 9- 1 (y) is a finite set {Xl, .. . , x r } each of whose points belongs to an
2. Inception and formation 25

(open) n-simplex of IKI. These simplices u are disjoint, and the restriction
of 9 to every u is a homeomorphism onto 9(U). Let p (resp., q) be the
number of simplices on which the restriction 910- preserves (resp., reverses)
the orientation, and consider the difference p - q. Then Brouwer shows
that the number p - q is the same for all y E fl off S. To that end, he
joins any two such y's by a polygonal contained in fl and moves y along:
p and q increase or decrease simultaneously by the same number when the
polygonal crosses S; hence p-q does not change. This happens because each
(n - 1)-simplex is a face of exactly two n-simplices of IKI. Consequently,
the integer d = p - q is well defined and is called the degree d(g) of g.
Step II: The general case. To deal with arbitrary non-simplicial mappings,
Brouwer introduces two constructions that have become fundamental in
combinatorial topology: barycentric subdivision of a simplicial complex
(subdivision of each simplex by taking its barycenter as a new vertex)
and simplicial approximation of continuous mappings. By means of these
refined geometric techniques, he defines the degree d(J) of an arbitrary
continuous mapping f : M -+ N through a good simplicial approximation
of f. In fact, he sees that any two close enough simplicial approximations
of f have the same degree as defined in Step 1. In fact, Brouwer shows that
close approximations are homotopic by a piecewise linear homotopy, which
implies the equality of degrees.

Let us make some pictures for this construction. Suppose we have the
following simplicial mapping:
PI

g
§l • §l

~ l~ l~
P4
VI '"
'"
VI

IKI 9= 1/Jgrp-l
• ILl
26 I. History

Here the localization 9 is affine on every 1-simplex of IKI; hence it is de-


termined by the images of the vertices: in this case we suppose

g(vo) = vo, g(Vl) = VI, g(V2) = V2,

g(V3) = vo, g(V4) = VI, g(V5) = V2·


Then take a point

and let g-l(y) = {Xl, X2}. In this situation, the restrictions

are homeomorphisms that preserve the orientation; hence the degree of J


is
d(f) = d(g) = p - q = 2 - 0 = 2.

Next, we look at the general case: a non-affine map like J(z) = z2. The
picture is the following:

§l
J §l

V2 ~ j~ ~
~

j~ 'ih

f = 'ljJgcp-l
IKI ILl
9
V4 Vo 112

Here 9 is the simplicial approximation of f, and we have

g(Vo) = vo, g(Vl) = VI, g(V2) = V2, g(V3) = V3,

g(V4)=VO, g(V5)=Vl, g(V6)=V2, g(V7)=V3.


2. Inception and formation 27

Computing as above, d(g) = 2, and we conclude d(f) = 2.

For his notion of degree, Brouwer obtains the following essential prop-
erties:
Proposition. The following statements hold:
(a) d(f) = 0 if f is not surjective.
(b) d(f) = ±1 if f is a homeomorphism.
(c) d(g 0 f) = d(g) . d(f) for the composition go f of two continuous
mappings f : M -+ N, g : N -+ P.
(d) d(iI) = d(h) if iI, h : M -+ N are homotopic.

Furthermore, Brouwer proves the following two important theorems:


Theorem. A continuous vector field tangent to an n-sphere of even di-
mension always has singular points (= zeros).
Theorem. A continuous mapping without fixed points from an n-sphere
into itself has degree -1 if n is even and degree +1 if n is odd.
Proof. The proof of the second result follows. First, the homotopy

tf(x) - (1 - t)x
F(t, x) = Iltf(x) _ (1 - t)xll

is well defined because f has no fixed point. Thus, the antipodal mapping
Fa is homotopic to f = F 1 , and thus Fa and f have the same degree, which
is ±1 according to the dimension. 0

From this, Brouwer deduces:


Corollary. (a) Every non-surjective continuous mapping of an n-sphere
must have some fixed point.
(b) Every continuous mapping of a sphere of even dimension that is
homotopic to the identity must have some fixed point.
(c) Every continuous mapping of a sphere of odd dimension that is ho-
motopic to the antipodal mapping must have some fixed point.
Proof. For (a), a non-surjective mapping has degree 0 '# ±1, and the pre-
ceding theorem applies. For (b), the degree of the given mapping is that
of the identity; hence it is 1 '# -1. As the dimension is even, again the
preceding theorem gives the conclusion. Assertion (c) is proven in the same
way, since the degree of the antipodal map in odd dimension is -1. 0
28 1. History

Another very important result due to Brouwer (and as mentioned be-


fore, revisited by Hadamard) is the following:
Brouwer Fixed Point Theorem. A continuous mapping from a closed
n-ball D'" into itself must have some fixed point.
Proof. The proof runs as follows. Given a continuous mapping f : D'" -t
D"', Brouwer identifies D'" with the upper hemisphere Wt of the n-sphere
§n and defines

~n -ti:l>~n :XHgx
g:i:l>
() = {f(X) if x E §n+,
f(Xl, ... , Xn , -Xn+l) if x ~ §+..
This mapping is continuous, but not surjective, hence has some fixed point
that must be a fixed point of f. D

With this powerful tool in hand, Brouwer proved in later articles many
other important results: the Invariance of Domain Theorem, the Jordan
Separation Theorem in arbitrary dimension, and various dimension proper-
ties. Moreover, in the paper [Brouwer 1912b], he defines the link coefficient
f(Kl, K 2) of two disjoint oriented compact manifolds K 1 , K2 in jRn, with
complementary dimensions hand n - h - 1. This link is the degree of the
mapping

that is,

We must mention here a contribution made 80 years earlier by Gauss,


who in [Gauss 1833] actually computed the link coefficient of two curves.
The Italian mathematician CARLO MIRANDA proved in [Miranda 1940]
that the Brouwer Fixed Point Theorem is equivalent to Poincare's Inter-
mediate Value Theorem of 1883 (1.1, p. 12). For this reason, some authors
call this theorem the Poincare-Miranda Theorem.

3. Accomplishment
This is the place to mention HEINZ HOPF (1894-1971). Hopf's interest in
degree theory came from the lectures given by ERHARD SCHMIDT (1876-
1959) at Breslau in 1917 and at Berlin in 1920 on the invariance of dimen-
sion and the proof of the Jordan Theorem in arbitrary dimension using
3. Accomplishment 29

Brouwer's degree. Hopf himself says in [Hopf 1966] that those results ap-
pealed to him deeply:
I was fascinated; this fascination-of the power of the method of the
mapping degree-has never left me since, but has influenced major parts of
my work.
Hopf's essential contribution to degree theory springs from his Ph.D.
thesis, which he defended at Berlin University in 1926. Among other as
important geometric questions, he completed Brouwer's results concerning
continuous mappings of closed oriented manifolds of the same dimension,
with special attention to the case when the target manifold is a sphere. We
explain it in the following paragraphs.

Let G and G' be two domains in the Euclidean n-space, and let f, g
be two continuous mappings from G into G'. Let a E G be an isolated
coincidence point of f and 9 (that is, f(a) = g(a) and this does not hold
true for any other point in a small enough neighborhood U C G of a). The
coincidence index of f and g at a is the degree of the continuous mapping

n-l f(x) - g(x)


<p: Sr(a) -+ § : x H IIf(x) _ g(x)II'

where Sr(a) is the spherical surface of center a and radius r > 0 contained
in U. By its local nature, this definition extends to mappings of arbitrary
manifolds. Modifying Brouwer's proof of the Fixed Point Theorem for
spherical surfaces, Hopf proved ([Hopf 1925]):
Theorem. Let f and 9 be continuous mappings from a compact oriented
manifold M of dimension n into the sphere §n, and denote by d, d' their
respective degrees. Then, if f and g have finitely many concidence points,
the sum of the coincidence indices of f and 9 at those points is ( -1 )nd + d'.
Brouwer had proved in [Brouwer 1912a] that continuous homotopic
mappings have the same degree and in [Brouwer 1912c] that continuous
mappings of §2 into itself with the same degree are homotopic. Hopf, using
the theorem above, proved in [Hopf 1926a] the most general form of this
fact:
Hopf Theorem. Two mappings from a compact oriented manifold M of
dimension n into the sphere §n that have the same degree are homotopic.
Later, in [Hopf 1933], Hopf extended this result to the case when the
source manifold M is a polyhedron. A further generalization to Banach
spaces was to be published by ERICH ROTHE in [Rothe 1936].
30 I. History

Now, the dual problem that arises in a natural way is the homotopic
classification of continuous mappings from a sphere into a topological space.
This led to the discovery of homotopy groups of higher order by EDUARD
tECH (1893-1960), who defined them in a very short paper presented
at the Zurich IeM, in 1932 [eech 1932]. However, these groups did not
attract much attention, as they are always commutative, in sharp con-
trast with the behavior of the fundamental group. But in 1935 WITOLD
HUREWICZ (1904-1956) rediscovered and studied systematically these ho-
motopy groups ([Hurewicz 1935a], [Hurewicz 1935b], [Hurewicz 1936a], and
[Hurewicz 1936b]). The two basic ingredients from which the theory devel-
ops are:

(1) the fundamental group (now first homotopy group), which Poincare
introduced in [Poincare 1895b], and

(2) the Hopf invariant, the means used by Hopf to distinguish infinitely
many homotopy classes of mappings §3 --+ §2. This he achieves by
combinatorial topology techniques in [Hopf 1931]. Some years later,
in [Hopf 1935], Hopf extends his result to mappings §2n-l --+ §n.

On the other hand, in [Poincare 1885a], Poincare had proved that if


a continuous vector field on a compact orient able surface of genus 9 has
finitely many critical points, the sum of their indices (Poincare indices) is
the topological invariant 2 - 2g, which is the Euler characteristic of the
surface. This was generalized by Brouwer in [Brouwer 1912a] to spherical
surfaces of arbitrary dimension. As we have said before, Hadamard an-
nounced without proof the same result for compact orientable manifolds of
arbitrary dimension [Hadamard 1910]. But it is Hopf who, finally, proved
in [Hopf 1926b] the following result:
Poincare-Hopf Theorem. Let M be a compact orientable manifold of
dimension n, and let X be a continuous vector field with finitely many crit-
ical points. Then the sum of the indices of those critical points is the Euler
characteristic x(M) of M.

But there is more to tell. In his path to these important results, Hopf
gained deeper insight into the topological meaning of the various invariants
involved. In fact, he analyzed Kronecker's work on the characteristic of
a system of several variables and defined the so-called integral curvature
of a closed oriented hypersurface M in the Euclidean (n + 1)-space. That
invariant is the degree of the Gauss mapping v of the hypersurface (v(x)
is the unitary vector perpendicular to the hypersurface at x). If M is de-
3. Accomplishment 31

scribed by an equation F = 0, the integral curvature of M is the Kronecker


characteristic of the system

( F, aaF
Xl
, ... , aaF ).
Xn+1

In the articles mentioned, Hopf proved:


Theorem. The integral curvature of a Jordan hypersurface in IR n+1 (one
that bounds a domain) coincides with the Euler characteristic of the hyper-
surface.
To refine this statement, Hopf introduced the models of a closed oriented
n-manifold. These are the hypersurfaces in IRn+1 that are homeomorphic
to M. Then, he proved:
Gauss-Bonnet Theorem in arbitrary dimension. Let M be a closed
oriented manifold of dimension n.

(1) If n is odd, the integral curvature of the models of M is not a topo-


logical invariant, not even for Jordan models.

(2) If n is even, the integral curvature of the models of M is a topological


invariant, namely, half the Euler characteristic of M.

If N is a model of M and K1, ... , Kn are its principal curvatures, then

1X(N) = vOlt§n) L K1 ... Kn dw .

This was obtained by WALTER FRANZ ANTON DYCK (1856-1934) in [Dyck


1888J and [Dyck 1890J.
In particular, we see that all models of even-dimensional manifolds have
even Euler characteristic, and thus we find manifolds which cannot be em-
bedded in co dimension 1:
Corollary. There are closed oriented n-manifolds which have no model in
IRn +1, even allowing self-intersections.
An example is the complex projective plane, which has Euler character-
istic 3 (as it can be computed with a vector field using the Poincare-Hopf
Theorem).

***
32 I. History

Brouwer-Kronecker degree theory and its applications were systemat-


ically developed by Hopf in the topology text [Aleksandrov-Hopf 1935]
written jointly with PAVEL SERGEEVICH ALEKSANDROV (1896-1982). The
subject is covered in the final three chapters, XII, XIII, and XIV, of that
book. The titles of the sections in those chapters give a detailed account
of the contents of the theory. We translate them as follows:

CHAPTER XII. The Brouwer degree. The Kronecker characteristic

§1. The order of a point with respect to a cycle. Recall results in link theory. The
Poincart1-Bohl theorem. The Rouche theorem. Homology of IR n \ {pl. The global degree.
Relation between order and degree. The winding number in the plane. The Kronecker
integral.
§2. The Kronecker characteristic. Local degree of mapping in IRn. The existence
theorem for a point. Application: The fundamental theorem of algebra. The characteristic
of a function system. The index of a point. The algebraic number at a point. The
local degree in IRn. Topological applications. The invariance theorem. The functional
determinant.
§3. Special theorems and applications. Relations among vector fields on IR n and map-
pings. Vector fields on the sphere. The Brouwer fixed point theorem on an n-dimensional
element. Vector fields on a spherical surface of even dimension: A fixed point theorem.
Symmetries on the spherical surface. Another fixed point theorem. Borsuk theorem on
antipodal mappings. Analytic corollary. Mappings from §n to IRn. The covering theorem
for the spherical surface §n.
§4. The degree of mappings between polyhedra. Definition of local degree. Contin-
uous deformations of mappings (homotopies). Cycle mappings. Cycle mappings on acyclic
irreducible closed complexes. Coincidence of the local and global degrees. Determination
of the global degree by homological invariants. Essential properties of mappings.
Appendix. Comment on the Brouwer link number as a characteristic. The
Gauss integral. The number of cuts as a degree. The link number as an order. The
Gauss integral.

CHAPTER XIII. Homotopy and mapping extension theorems

§1. More on the Kronecker existence theorem. An extension problem for mappings
into IRn. Reduction to mappings into §n. Elementary lemma on extension and homotopy
of mappings. Solution to the extension problem for a simplex. Reduction of Theorem II
to a lemma. Algebraic part of the proof of the lemma. Geometric part of the proof of the
lemma. Criterion for the local essentiality of mappings on lRTl..
§2. Mappings from n-dimensional polyhedra into the spherical n-dimensional
surface. Equivalence between homology and homotopy. Listing of the types of mappings
in the simplest case. Essential mappings. Resolution theorem.
§3. Mappings from n-dimensional polyhedra into the circle. Mappings from
the n-dimensional sphere into the circle. The extension theorem. Types of mappings.
Essentiality.
§4. Characterization of the identity and polyhedra borders by deformation
properties. A deformation theorem from a polyhedron into itself. The essentiality of a
polyhedron into itself. The particular case of polygons. Characterization of the identity.
Characterization of the border. Stability of polyhedra. Examples.
Appendix. Mappings that are homologous but not homotopic.
3. Accomplishment 33

CHAPTER XIV. Fixed points

§1. A fixed point existence theorem. Fixed simplex. A generalization of the Poincare-
Euler formula. The Lefschetz number of a continuous mapping of a polyhedron into itself.
Fixed point existence theorem. Examples. Remarks.
§2. The index of fixed points. The index of a zero of a vector field and of a singular
point of an oriented field. The index of a fixed point. Index properties. Normal fixed
points. Fixed points of affine mappings. Topological invariance of the index of fixed
points. The invariance of the index at a singular point of an oriented field.
§3. Algebraic number of fixed points of a continuous mapping from a poly-
hedron into itself. The algebraic number of regular fixed points. General fixed points
theorem. Regular fixed points of simplicial mappings. Reduction to the approximation
theorem. Proof of the approximation theorem. Remarks on the concept of number of fixed
points.
§4. Oriented fields on closed manifolds. Preliminaries on smooth manifolds. Oriented
fields and their singularities in manifolds.

***
After this, the next step in degree theory was taken by JEAN LERAY
(1906-1998) and JULIUSZ PAWEL SCHAUDER (1899-1943), who developed
degree theory for completely continuous mappings in Banach spaces. They
published their results in 1934 in the paper [Leray-Schauder 1934]. Previ-
ously, a summary of the paper had been presented at the Paris Academy
of Sciences [Leray-Schauder 1933] on July 10, 1933. This summary starts
as follows:

Let £ be an abstract space, normed, linear, and complete (in the sense
of Mr. Banach [Banach 1922]); let W be an open set of £ (whose boundary
we denote by W', and its closure by W = W + W'); finally, let

y = x - F(x) = 4>(x)

be a functional transformation defined on W. We do not suppose F(x) lin-


ear, but only completely continuous (or compact, vollstetig) (that is, F(x)
is continuous and transforms every bounded set into a relatively compact
one); furthermore, F(x) takes values only in £. We have been able to
define, after the famous work by Mr. Brouwer [Brouwer 1912a], the topo-
logical degree d[4>, W, b] of the transformation 4> at a point b off 4>(W'), in
such a way that the well-known properties of this degree are still valid. We
describe briefly our definition.
The Leray-Schauder construction is this. Let us suppose the point b is
the origin 0 E £, and set

h = min{d(O,4>(x)): x E Wi}.
34 1. History

Consider a new functional transformation Fh(X) at a distance < h from


F(x) and such that all its values belong to a common linear subset M of £
of finite dimension; denote by W M the intersection of W with that subset
M. Let <Ph(X) = x - Fh(X). We set

where the right-hand side degree comes from Brouwer's theory.


Then, the authors prove that the value d(<p, W,O) does not depend on
the choice of <Ph and W M, after which they develop the main properties of
this degree. They introduce the following:
Definition. The degree at 0 of the transformation

(1) y=x-F(x) = <p(x)


is called the total index of the solutions in W of the equation

(2) x-F(x) =0.

With this terminology, they prove:


Theorem. Suppose that equation (2) depends continuously on a parameter
k varying in a segment K of the real line. Let il be an open bounded set of
the space £ x K such that (i) F is defined on il and (ii) the boundary il' of
il does not contain solutions to (1). Then, the total index of the solutions
in the interior of il is the same for all values of k.
Thus, we see that the existence of at least a solution to (2) is guaranteed
for every value k as soon as one knows a value ko for which the total index
is not zero. The authors conclude by stressing that their theorem has
a wide range of applications. In particular, it applies to limit (boundary)
problems concerning elliptic second-order partial differential equations, and
they can use it to generalize the well-known existence theorems due to
SERGEI NATANOVICH BERNSTEIN (1880-1968).
These important applications to differential equations of the Leray-
Schauder methods lead to the consideration of more general operators,
in particular linear and non-linear Fredholm or Noether operators. The
first investigations in this line are due to RENATO CACCIOPPOLI, who in
[Caccioppoli 1936] obtained a mod 2 degree theory for continuously differ-
entiable Fredholm mappings of index O.
4. Renaissance and reformation 35

4. Renaissance and reformation


After the publication of Brouwer's fundamental papers, much effort was
devoted to establishing the main properties of the degree of a continuous
mapping, and in particular the Fixed Point Theorem, by analytic methods
that did not involve the concepts of combinatorial topology. It is clear that
the first satisfactory approach to this question was Hadamard's use of the
Kronecker integral. The second attempt is based on the paper [Sard 1942]
by ARTHUR SARD (1909-1980). There, we find the famous:
Sard Theorem. Let G be an open set in ]Rm, and let f : G -t ]Rn be a CP
mapping {p ~ I}, with p ~ m - n + 1. Let Cf stand for the set of critical
values of f, that is, y = f(x) E Cf when the Jacobian matrix of f at x has
rank < min(m, n) {which is the maximum possible rank}. Then, Cf has
Lebesgue measure zero.
In particular, Cf has empty interior, hence dense complement, in ]Rn.
The final conclusion comes from the fact that a subset of]Rn whose
Lebesgue measure is zero does not contain any open set, a result proved
previously by ARTHUR BARTON BROWN in [Brown 1935]. Due to this, the
theorem above is often called the Bard-Brown Theorem.
Using this theorem and the classical Weierstrass Theorem on polyno-
mial uniform approximation of continuous functions on compact sets in
Euclidean spaces, in 1951 MITIO NAGUMO constructed Brouwer's degree
theory by elementary analytic means. He published the results in [Nagumo
1951a]; the journal had received the paper on March 6, 1950. In the intro-
duction, the author says:
This paper establishes a theory of degree of mapping for open sets in
a Euclidean space of finite dimension, based on the theory of infinitesimal
analysis, which is free of the notion of simplicial mapping. Although the
results are not new, I hope in this way to make it possible to incorporate
the theory of degree of mapping into a course in infinitesimal analysis.
The basic ideas of Nagumo's construction are these:
Btep I: Let G be an open bounded set in ]Rn, f : G -t ]Rn a C1 mapping,
and a E ]Rn \ (J (G \ G) U Cf). Then, by the Inverse Function Theorem,
the derivative df(x) is a linear isomorphism for every x E f-l(a), and con-
sequently f-l(a) is a finite set, say f-l(a) = {Xl, ... , xp}. Thus, Nagumo
36 1. History

can define

Step II: Let G be an open bounded set in IRn, f : G ~ IRn a continuous


mapping, and a E IRn \ f (G \ G). Then d(g, G, a) has the same value for all
C2 mappings 9 : G ~ IRn such that

(1) Ig(x) - f(x)1 < dist (a, f(G \ G)) for all x E G, and

(2) a E IR n \ e g.

Notice here that condition (1) implies a ~ g(G \ G); hence Step I applies
to g, and Nagumo can also define the degree of f at a by setting

d(j, G, a) = d(g, G, a)
for any 9 as above.
We remark that Step II is where Nagumo used the Sard-Brown and
Weierstrass Theorems. On the other hand, it is worth stressing that the
degree d(j, G, a) is the Hadamard index of flaG, in case BG = G \ G is a
hypersurface.
This same method leads to the construction of degree for continuous
mappings of boundaryless compact oriented manifolds, once one extends
the Sard and Weierstrass Theorems to those manifolds. It is easy to see
that the degree obtained this way coincides with Brouwer's.

Nagumo also wrote a second important article, [Nagumo 1951bJ, where


we find the first axiomatic approach to degree theory. This second arti-
cle was also received by the journal on March 6, 1950, and published in
the same volume right after the first one. In the introduction, the author
explains that the goal of the paper is to complete the details of Leray-
Schauder's construction of degree and that following their ideas, he will
describe a degree theory for locally convex topological spaces and prove the
Invariance of Domain Theorem for locally convex complete linear spaces.
From this paper, we extract paragraph 1.1, relevant for the axiomatic de-
velopement of degree theory. Nagumo starts:
First we shall explain the notion of degree of mapping in a finite dimen-
sional Euclidean space.
4. Renaissance and reformation 37

And he presents the following existence statement:


Nagumo axiomatization. Let G be an open set in Rm and let f : G --t
Rm be such that f(x) - x is bounded for x E G. Let a fJ. f(G \ G). Then
there will be determined an integer d(f, G, a), called the degree of mapping
of G at a by f, with the following properties:

(i) d(Ida , G, a) = 1 for a E G; d(Id a, G, a) = 0 for a fJ. G.


(ii) If d(f, G, a) i= 0, then there is an x E G such that f(x) = a.
(iii) If G ::) Ui G i , G ::) Ui Gi , for finitely many disjoint open sets Gi, and
a fJ. f(G i \ Gd for any i, then

d(f,G,a) = 'Ld(Ii,G,a).
i

(iv) If h(x) - x is a bounded continuous function of (t,x) for 0:::; t:::; 1,


x E G, if a(t) ERn is continuous, and if a(t) fJ. h(G \ G) for 0:::; t :::;
1, then d(h, G, a(t)) is constant for 0 :::; t :::; l.

(v) Let X be the set of all roots of the equation f(x) = a in G, and let
Go be any open set such that X c Go c G. Then

d(f, G, a) = d(f, Go, a).

Then, Nagumo remarks that (v) follows from (ii) and (iii) and concludes:
The existence and uniqueness of d(f, G, a) satysfying the above condi-
tions can be verified, if we use simplicial mappings for approximations of f
(at first for bounded G and then for general G). But I may refer to [N agumo
1951a] in which the existence of d(f, G, a) is given, based on infinitesimal
analysis but free from the notion of simplicial mapping.
Only one objection can be made here: Nagumo does not prove unique-
ness!

***
We close this section by recalling two additional ways to define the
Brouwer-Kronecker degree.
De Rham cohomology. On one hand, GEORGES DE RHAM (1903-1990)
published the book [de Rham 1955], where he developed homology and
38 I. History

degree theory from a completely different viewpoint. He used flows in


connection with the theory of distributions due to LAURENT SCHWARTZ
(1915-2002). We can describe this method very roughly as follows.
Every differentiable mapping f : M -t N of boundaryless compact
connected oriented n-manifolds induces in (de Rham) cohomology a linear
mapping f* : Hn(N, JR) -t Hn(M, JR). Since Hn( ., JR) == JR, that linear
mapping is just multiplication times some number, but this number is an
integer d, which we call the degree of f. Among other things, two important
results are behind this approach: Stokes' Theorem and the Change of
Variables Formula for Integrals.
Heinz Integral Formula. On the other hand, and related to the above
method, in 1959 ERHARD HEINZ presented a more elementary construction,
which appears in [Heinz 1959J and runs as follows.
Let y = y(x) be a C1 mapping defined on a bounded open set fl c JRm,
which is continuous on fl. Furthermore, assume we are given a point z E
JRm such that y(x) i- z for all x E fl \ fl, and let lP(r) be a real-valued
function such that the following hold:

(1) IP( 1") is continuous on the interval 0 ~ r < +00 and vanishes on a
neighborhood of 1" = 0 and on c ~ r < +00, where

0< c < min{y(x) - z : x E fl \ fl}.

(2) fIRm 1P(lxl)dx = 1.

Then the Brouwer degree d(y(x), fl, z) is uniquely defined by

d(y(x), fl, z) = In 1P(ly(x) - zl)J(y(x))dx.

5. Axiomatization
The development of degree theory was fully completed in 1971, with the
axiomatic characterization of the Brouwer-Kronecker topological degree.
Such a characterization was obtained by LUTZ FUHRER in his Ph.D. thesis
[Fuhrer 1971 J, presented at the Freie Universitat Berlin. These results were
published in the paper [Fuhrer 1972J, received by the journal on October
28, 1971. Let us also mention that Fuhrer quoted Nagumo.
5. Axiomatization 39

However, the axiomatic characterization of degree is often attributed to


HERBERT AMANN and STANLEY A. WEISS. In fact, in the paper [Amann-
Weiss 1973] (received by the journal on August 21, 1972) they characterize
axiomatically not only the Brouwer-Kronecker degree, but also the Leray-
Schauder degree. The authors conclude their introduction with the follow-
ing acknowledgement:
After having finished the first draft of this paper the authors learned of
the thesis of Fiihrer [Flihrer 1971] in which the uniqueness of the Brouwer
degree has been proved also.
Thus, priorities are clear: the axiomatization of the Brouwer-Kronecker
degree is due to Flihrer, that of the Leray-Schauder degree to Amann-Weiss.

***
Let us start with the following:
Fiihrer Characterization. There exists a unique mapping

Dc jRn bounded open, }


d : { (J, D, y): f: D -+ jRn continuous, ----+ Z
y E jRn \ f(8D)

such that we have the following:

(1) Homotopy invariance: For every bounded open set D C jRnand all
continuous mappings F: [0,1] x D -+ jRn and"( : [0,1] -+ jRn such
that
"((t) E jRn \ F( {t} x 8D) for 0:S t :S 1,
the following formula holds:

d(F(t, .), D, "((t)) = d(F(O, .), D, "((0)) for O:S t :S 1.

(2) Normality: For every bounded open set D C jRn and every point
pED,
d(Id]), D,p) = 1.
(3) Additivity: For every bounded open set D C jRn, every disjoint union
D = Dl UD2 of two open sets, every continuous mapping f : D -+ jRn,
and every point p E jRn \ f(8D U 8Dl U 8D2),

d(J, D,p) = d(J, D1,p) + d(J' D2,p).


40 I. History

(4) Existence of solutions: For every bounded open set D c IR n , every


continuous mapping f : D -+ IR n , and every point p E IRn \ f(aD),
such that d(f, D,p) i= 0, the equation f(x) = p has some solution in
D.

This is the full description of the Brouwer-Kronecker degree. This on


one hand has the virtue that the theory becomes clearer and systematic,
but on the other hand, and this is very important, this also shows that
all different definitions of degree invented by the 1960s coincide, no matter
which topological or analytical means were involved.

As often happens, once a collection of complete axioms is found, there is


a search for simplicity that aims to distill the collection as much as possible.
In this respect, it is worth quoting here the following result obtained by
KLAUS DEIMLING in [Deimling 1985]:
Deimling Characterization. There exists a unique mapping
Dc IRn bounded open,
{
d: (f, D, y): f: D -+ IRn continuous,
y E IRn \ f(aD)
such that we have the following:

(1) Normality: For every bounded open set D C IRn and every point
pED,
d(Idl), D,p) = 1.

(2) Additivity: For every bounded open set D C IRn , every pair of two
disjoint open sets Dl, D2 cD, every continuous mapping f : D -+
IR n , and every point p rt f(D \ Dl U D2),

d(f, D,p) = d(f, D1,p) + d(f, D 2,p).


(3) Homotopy invariance: For every bounded open set D C IRn and all
continuous mappings H: [0, 1] x D -+ IRn and y : [0, 1] -+ IRn such
that
y(t) rt H({t} x aD) for 0 ~ t ~ 1,
the following formula holds:

d(H(t,·),D,y(t))=d(H(O,·),D,y(O)) for O~t~1.

***
5. Axiomatization 41

Next we describe the axiomatization of the Leray-Schauder degree ac-


cording to Amann-Weiss.
Let E be a separated topological linear space, and consider a family W
of open subsets including 0 (but not only 0). To each S1 E W, attach a set
M (S1) of continuous mappings S1 -t E. We equip this set M (S1) with the
uniform convergency topology.
Definition 1. The family
M(W) = {M(S1) : S1 E W}
is called an admisible class of mappings on E when the following properties
hold true:

(1) Idn E M(S1) for every non-empty open set S1 E W.


(2) For every pair of two open sets S11 C S1 in Wand every f in M(S1),
the restriction flnl belongs to M(S1I ).

Now, fix such an admissible class M(W) for the given family W of open
sets of E. For every S1 E W, consider the following subspace of M(S1):
Mo(S1) = {f E M(S1) : 0 ~ f(aS1)},
where as usual aS1 = S1 \ S1 stands for the boundary of S1. Then:
Definition 2. A topological degree d for M(W) is a family of mappings
d = {d(·, S1) : Mo(S1) -t Z: S1 E W}
for which the following conditions hold true:

(1) Normality: d(Idn , S1) = 1 for every S1 E W with 0 E S1.


(2) Additivity: For every non-empty open set S1 E W, every pair of two
disjoint open subsets of S1, S11 , S12 E W, and every f E M (S1) such
that 0 ~ f(S1 \ S11 U S12 ),

d(f, S1) = d(flnl' S1t) + d(fln2' S12 ).

(3) Homotopy invariance: For every non-empty open set S1 E Wand


every continuous mapping h : [0,1] -t Mo(S1),
d(h(t), S1) = d(h(O), S1) for 0 ~ t ~ 1.
42 1. History

Once the setting is fixed in this way, the authors prove their main result:
Theorem. Let E be a locally convex linear space, and let W be either (i)
the family of all open sets of E or (ii) the family of all bounded open sets of
E. Then, there exists a unique topological degree for the admissible class
M(W) = {K(.!1) : .!1 E W},

where K(.!1) consists of all continuous mappings I : .!1 -t E such that the
image (Idn - f) (.!1) is relatively compact (f is a compact vector field).
Of course, this implies the uniqueness of both the Brouwer-Kronecker
and the Leray-Schauder degrees.

6. Further developments
There are several lines along which generalizations have been developed,
with various aims and scope. We discuss here: (i) the case of spaces of
infinite dimension, (ii) the case when source and targent have different
dimensions, and (iii) the equivariant case.

***
Degree theory in infinite dimension. This line of research comes from
Leray-Schauder theory. Its main purpose is to construct degrees for vector
fields other than the compact ones, i.e., Fredholm, monotonous, contrac-
tive, ... , and to find the corresponding axiomatizations. The resulting the-
ories apply to problems on partial differential equations and bifurcation in
functional equations. As we mentioned before (1.3, p. 34), this was started
by Caccioppoli in 1936. Years later, the idea was rediscovered and pre-
sented in a more rigorous general way by STEPHEN SMALE in [Smale 1965],
making use of the non-oriented cobordism rings invented by RENE THOM
in his outstanding foundational paper [Thorn 1954]. Smale's definition can
be summarized as follows.
Let I : M -t V be a proper Fredholm mapping of index p ~ 0 and
class q > p+ 1. Then the non-oriented cobordism class of the inverse image
f-l(a) of a regular value a is a well-defined invariant "(U), which vanishes
if f is not surjective. In case the index p is zero, then 1-1 (a) is a finite set
and "(U) coincides with Caccioppoli's mod 2 degree.
When orientations are taken into account for index 0 Fredholm opera-
tors, the result is a degree theory with integral values, a construction made
6. Further developments 43

by K. DAVID ELWORTHYand ANTHONY J. TROMBA in [Elworthy-Tromba


1970a] and [Elworthy-Tromba 1970b]. In the first of these two references,
the authors present an oriented degree theory for proper Fredholm map-
pings of index n and class r = n + 2 using framed cobordism as introduced
by LEV SEMENOVICH PONTRYAGIN in [Pontryagin 1955] (a gem of topol-
ogy).
Then, in [Nirenberg 1971], LOUIS NIRENBERG produced a generalized
topological degree theory for compact perturbations of Fredholm opera-
tors, using stable homotopy groups and their analogous version in infinite
dimension. These results are extended by E. NORMAN DANCER in [Dancer
1983]. Another interesting contribution to this topic is due to JORGE IZE,
who used cohomotopy groups in [Ize 1981].

***
Mapping degree for source and target of distinct dimensions. From
the preceding discussion, we see how bifurcation theory motivates the in-
troduction of oriented degree theories for Fredholm mappings of positive
index and therefore brings in the notion of topological degree for map-
pings between spaces of different dimensions. In such theories the so-called
degree is not an integer anymore, but some homotopy class in a suitable
homotopy group (of a sphere, because spheres are compactifications of Eu-
clidean spaces). An important development of this theory is presented by
KAZIMIERZ GEBA, IVAR MASSAB(), and ALFONSO VIGNOLI in [Geba et al.
1986].
The goal is to construct a generalized degree (the Geba-Massabo- Vignoli
degree) for continuous mappings f : U -+ ]Rn, where U is a bounded open
subset of ]Rm, with m ~ n, under the assumption that f does not vanish on
au = U\U. The authors support their theory by exploring the m = nease,
that is, by reformulating the Brouwer-Kronecker degree in a way suitable
for generalization. Indeed, let f be given as above, with m = n, and
denote by fo : au -+ ]Rn \ {O} the restriction of f to au. We look at ]Rn
and ]Rn \ {O} inside the Alexandroff compactijication (]Rn)* of ]Rn. Since
(]Rn)* \ {O} is homeomorphic to ]Rn, by the Tietze Extension Theorem, fo
extends to a continuous mapping foe : (]Rn)* \ U -+ (]Rn)* \ {O}. Thus one
gets a continuous mapping
44 I. History

Such an extension is called admissible, but note that it is not unique. Next,
we consider a homeomorphism an : (ffi.n )* --+ §n such that

an(oo) = (1,0, ... ,0),


an((ffi.nr+) = {x E §n : xn+l ~ a},
an((ffi.n):..) = {x E §n : Xn+l ~ a},
an+l(x) = an(x) for all x E ffi. n.

It can then be shown that the homotopy class of an 0 f* 0 a;;1 does not
depend on the admissible extension f*, and thus we have a well-defined
element [an 0 f* 0 a;;1] in the n-th homotopy group 7rn (§n). Next, via an
algebraic isomorphism 'Pn : 7rn(§n) --+ Z, set

Once d* is thus constructed, it is checked that the axioms of degree theory


hold true, and by the axiomatic characterization (I.5, p. 38), d* is indeed
the Brouwer-Kronecker degree.
Now, this construction can be mimicked for arbitrary m ~ n to obtain
a generalized degree:

which, for m > n, is not an integer any more. Moreover, we see why the
case m < n was neglected: in that case 7rm(§n) = 0.
In the article mentioned, the authors prove for this degree the basic
properties, namely:

(1) Homotopy invariance. If h : [0,1] x (U, aU) --+ (ffi.n , ffi.n \ {o}) is a
continuous mapping, then d*(ht, U) is well defined for all t and does
not depend on t.

(2) Excision. Let f : (U, aU) --+ (ffi.n , ffi.n \ {o}) be a continuous mapping.
Then, for every open set V C U such that f has no zeros in U \ V,
we have d*(f, V) = d*(f, U).

(3) Existence of solutions. Let f : (U, aU) --+ (ffi.n , ffi.n \ {o}) be a contin-
°
uous mapping with d*(f, U) =1= E 7rm(§n). Then, there is an x E U
such that f(x) = 0.

(4) Suspension. Let U be an open subset of ffi.m+l and let

f : (U, aU) --+ (ffi.n+l, ffi.n+l \ {o})


6. Further developments 45

be a continuous mapping such that

f(U n !R.~+1) C !R.++1, f(U n !R.~+l) C !R.~+1 .


Then, setting Uo = Un!R. m == un (!R.m x {O}) and fo = fluo' we have
fo: (Uo,aUo) -+ (!R.n,!R. n \ {O}) and

d*(f, U) = E(d*(fo, Uo)),


where E : 7rm(§n) -+ 7rm +l (§n+1) is the suspension homomorphism
(an isomorphismform < 2n-1 and an epimorphismform = 2n-1).

(5) Additivity. Let f : (U, aU) -+ (!R. n , !R.n\ {O}) be a continuous mapping,
and let U1 , U2 CUbe two open disjoint sets such that f has no zero
in U \ U1 U U2. Then d*(f, U) = d*(f, Ul) + d*(f, U2), whenever
m - n:S n - 4.

Later FRANCISCO ROMERO RUIZ DEL PORTAL showed in his Ph.D.


thesis [Ruiz del Portal 1991] that additivity also holds for m - n :S n - 2,
and this is definitive: there is a counterexample for m = 2n + 1. This
appeared in [Ruiz del Portal 1992]. Another counterexample was published
afterwards by Ize, Massabo, and Vignoli in [Ize et al. 1992]' a paper that
deals with equivariant degree as explained below.
To conclude, it must be noted that one major problem of this theory is
the computation of the homotopy groups of the spheres, a question which
is wide open today.

***
Equivariant degree theory. The purpose here is to define a suitable
topological degree for mappings that are invariant under the action of a
Lie group G on the given spaces. The case most studied is that of G = §l,
that is, the so-called §l-equivariant topological degree.
Let us recall that Poincare used what later would be called the Brouwer-
Kronecker degree to study the critical points of a differential equation. But,
as is well known, other very important elements for the understanding of
differential equations are periodic orbits. It was to count their number that
in 1965 F. BROCK FULLER introduced an invariant of flows that today
we call the Fuller index. The definition and properties of this index are
given in detail in [Fuller 1967]. A careful analysis of the constructions
behind the Fuller index and the generalized Geba-MassabO-Vignoli degree
46 1. History

for m = n+lled the above-mentioned Geba and GRZEGORZ DYLAWERSKI,


JERZY JODEL, and WAC LAW MARZANTOWICZ, from Gdansk University, to
define in a preprint in 1987 (later published as [Dylawerski et al. 1991])
a new homotopy invariant for §l-equivariant continuous mappings, which
they called §l-degree. Let us describe this briefly.
Let p be a finite representation of §l, that is, a continuous homomor-
phism p : §l -+ GL(V) into the linear group GL(V) of a real linear space
V of finite dimension; p determines an §l-action on V by (g, v) H- p(g)(v).
For such a pair (V, p), we say the following:

(1) A set X C V x JR is invariant if


(p(g)(v),..\) E X for all g E §\ (v,..\) EX.

(2) A continuous mapping f : X -+ V with invariant domain X is an


§l-mapping if

f(p(g)(v),..\) = p(g)(j(v,..\)) for all g E §l, (v,..\) EX.

We will denote by 21. the abelian group of all finite sequences 0: = (O:r)r20,
with 0:0 E Z2 and O:r E Z for r ;::: 1 (sum defined componentwise).
With this terminology fixed, the authors prove:
Theorem. Let (V, p) run through finite representations of §l, il through
the family of all bounded, invariant open subsets of V x JR, and f : X -+ V
through §l-mappings such that X is invariant, il c X, and f(ail) c
V \ {o}. Then there exists an 2l-valued function Deg(j, il), called the §l_
degree, satisfying the following conditions:

(a) If Deg(j, il) =1= 0, then f-l(O) n il =1= 0.


(b) If ilo c il is open and invariant and f-l(O)nil c ilo, then Deg(j, il)
= Deg(j, ilo).

(c) If ill, il2 are two open invariant subsets of il such that ill n il2 = 0
and f-l(O)nil C ill Uil2, then Deg(j, il) = Deg(j, ill) + Deg(j, il2)'

(d) If h: (il x [0, l],ail x [0,1]) -+ (V, V \ {O}) is an §l-homotopy, then


Deg(ho, il) = Deg(hl' il).
(e) Suppose (W, 1J) is another representation of §1 and let U be an open
bounded, invariant subset of W such that 0 E U. Define F : U x il -+
W x V by F(x, y) = (x, f(y)). Then Deg(F, U x il) = Deg(j, il).
6. Further developments 47

The construction of Deg(j, n) begins with two particular cases. The


first case is when we have the trivial representation V = lRn of §1. Then the
§1-mappings in the theorem are the continuous mappings 1 : (n, an) -t
(lR n , lRn \ {O}), where n is a bounded open subset of lRn+1. In this case
we have a homomorphism E : 1Tn+1(§n) -t Z2 (suspension) that is an
isomorphism for n ~ 3, and the authors define

Deg(j, n) = E(d*(j, n)) E Z2,


where d* is the Geba-Massabo-Vignoli generalized degree. In this case, the
proof of the theorem above follows from the results in [Geba et al. 1986].
Secondly, suppose there exists a point a E n such that
1-1(0) = {p(g)(a) : 9 E §1};

that is, 1-1(0) is an orbit of the action of §1 over V. Then the set

§1 *a = {g E §1 : p(g) (a) = a}

is finite, hence a subgroup of §1 consisting of k-th roots of unit. We can


furthermore assume without loss of generality that (V, p) is orthogonal (with
respect to some inner product on V) and decompose

V = W E9 Wl., W = {x E V: p(g)(x) = x}.

Then if in that decomposition 1 takes the form 1(x,y,>.) = (h(x,y,>.),y),


the authors define

Deg(j,n) = (a r ), ar = {deg(j, D) for r = k,


o otherwise,

where deg(j, D) is the Brouwer-Kronecker degree and D is a closed disc in


the linear space W, contained in nn(W x lR), transversal to §1 *a = 1- 1(0),
and oriented so that it can be identified with the unit disc in W.
In the general case, the construction of Deg requires (i) the classical
theorem that gives the full classification of all finite-dimensional representa-
tions of §1 ([Adams 1969]) and (ii) a quite non-trivial homotopy argument.
Another version of this degree for mappings defined on spheres has
been studied using cohomological obstruction theory by Ize, Massabo, and
Vignoli, in [Ize et al. 1986]. Afterwards, the same authors presented in
the two papers [Ize et al. 1989] and [Ize et al. 1992] an equivariant degree
theory for mappings defined on the closure of an arbitrary open set of the
48 1. History

ambient space and replaced the circle group §1 by an arbitrary compact


Lie group. In addition, they avoided the use of obstruction theory. This
is different, in the non-equivariant setting, from the Geba-Massabo-Vignoli
degree described earlier.
Namely, let U be a bounded open subset of ]Rm and let f : U -+ ]Rn
(m 2: n) be a continuous mapping such that f (x) i= 0 for x E au. Let
1: B -+ ]Rn be a continuous extension of f to a closed ball D containing
U. Let V be a bounded open neighborhood of au with f(x) i= 0 for x E V.
Consider a Uryshon function cp : D -+ [0,1] which is == 1 off U U V and == 0
on U. Define a mapping

F: [0,1] x D -+ ]Rn+1 : (t, x) f-+ F(t, x) = (2t + 2cp(x) - 1, !(x)).

It is easy to see that F(t, x) = 0 only if x E U, f(x) = 0, and t = ~.


Thus F maps a([O, 1] x D) into ]Rn+l \ {O}, which defines an element of the
homotopy group 7rm (§n): this is the generalized degree of f with respect to
U. When the action of a compact Lie group G is present, this construction
extends without difficulty, and the equivariant degree is an element of the
equivariant homotopy group 7r~(§n). Then, in [Ize et al. 1992] the authors
proved that this formulation for G = §1 is the §l-degree of [Dylawerski et
al. 1991].
All of this is revised in systematic form in the book [Ize-Vignoli 2003].
Also, we mention that Dancer defined an §l-degree for gradient map-
pings in [Dancer 1985]. In this respect, the paper [Geba et al. 1990], by
Geba, Massabo, and Vignoli, is remarkable. Later, SLAWOMIR RYBICKI
discussed an §l-degree for orthogonal mappings (which include gradient
mappings, [Rybicki 1994]), drawing upon the earlier work in [Dylawerski et
al. 1991]. To end these comments, let us add that Dancer, Geba, and Ry-
bicki obtained in [Dancer et al. 2005] a complete classification of equivari-
ant gradient mappings up to homotopy, and the corresponding equivariant
homotopy classes can be seen as equivariant degrees.
Chapter II

Manifolds
Here we describe the objects of our theory: manifolds and mappings. We will
be dealing with differentiable manifolds: our methods are those of differential
topology. The general definitions and some basic facts are gathered in §§1-2. In
particular, we stress the distinction between differentiable and smooth structures.
In §3 we state the essential Sard-Brown Theorem for differentiable mappings but
prove it only for smooth mappings. This brings in the subtleties concerning finite
or infinite differentiability, but we have chosen a most dramatic approach: to use
the fact that every differentiable structure is in fact smooth. In §4 we discuss in
depth the existence and two different constructions of tubular neighborhoods and
dijJerentiable retractions, both of which will be required later. Then, in §5 we
prove a key result for degree theory: differentiable mappings that are homotopic
are dijJerentiably homotopic. This is essential, because even for smooth manifolds,
we use our methods to study continuous mappings, neither smooth nor even differ-
entiable. In §6 we consider a special type of homotopies: the so-called dijJeotopies,
often needed to move points in manifolds at will. Finally, in §7 we recall the basics
of orientation, which playa crucial role in degree theory.

1. Differentiable mappings
In this section we recall the basic notions concerning differentiability. In
particular we discuss partitions of unity and bump functions.

(1.1) Differentiable mappings. Let U c IRP be an open set and let


r = 1, ... ,00. A differentiable cr function f : U -t IR is a function whose
partial derivatives ax' ~~ ~x' exist and are continuous for all k :S r. We use
'1 'k
the terminology smooth when r = 00. A mapping f : X -t Y of arbitrary
sets X c IRP and Y c IRq is Cr if for every point x E X there are Cr func-
tions Ii : U -t IR, 1 :S j :S q, defined on an open neighborhood U of x in
IRP such that J = (Jl, ... , Jq) coincides with f on U n X j we say that J is
a local Cr extension of f.

- 49
50 II. Manifolds

Notice how the extension f depends on the point x and that this is
a local notion. This calls in the main globalization tool in our context:
partitions of unity.
Proposition 1.2. Let U = {Ui : i E I} be an open covering of an open set
U C ]RP. Then, there is a smooth partition of unity {(h : i E I} for U, that
is, a family of smooth functions Oi : U ---+ [0, 1] such that:

(i) Each x E U has a neighborhood W on which all but finitely many Oi 's
vanish (local finiteness), and Li Oi == 1.

(ii) Each function Oi vanishes off Ui, namely {x E U : Oi(X) i- O} CUi.


This result is quite elementary in nature, and various proofs can be
found in many textbooks. We include a condensed version here for the
sake of the reader.

Proof. Consider a cover of U by compact sets L k , k 2:: 0 such that Lk C


Int(Lk+l) and Lo is a singleton, so that the compact sets
K = Lk-:-l \ Int(Lk) C U
also cover U.
Fix k. For every a E K there are (1) an open neighborhood D C
Int(Lk+2) \ Lk-l with D C Ui(k) for some i(k) and (2) a diffeomorphism
x: D ---+ ]RP with x(a) = O. Now define on D the smooth function
f(2 - Ilx11 2 )
1]= f(2 - IIx11 2 ) + f(ll xl1 2 -
1)
where f(t) = exp( -1ft) for t > 0 and f(t) = 0 for t ~ O. This function is
always 2:: 0 and it is == 1 on an open neighborhood BcD of a and extends
to U by 0 off D. In particular, 1] vanishes on Lk-l.

Graph of rJ (in ~P+1) Domain of rJ (in ~P)


1. Differentiable mappings 51

Next, since the open sets B cover the compact K, finitely many, call
them BR, already cover K, and we have the associated functions 'fJkf. We
have thus constructed for the fixed k the finite familly of non-negative
smooth functions {'fJkt}e, each vanishing off an open set Ui(ke); in fact
{'fJkR -=J O} c Ui(kR)'
Now let x E U, and pick the first ko with x E Lko' By construction,
each 'fJkR vanishes on Lk-l, which for k ~ ko + 2 contains the open set
W = Int(Lko+d, which contains x. Thus, the sum L.kR 'fJkR is actually finite
on W. Moreover, again by construction, 'fJkf(X) -=J 0 for some k,R.; hence
that sum is > O. This shows that any sum of the functions 'fJkR is a well-
defined smooth function, and the sum h of all of them is always > O. From
these remarks, we define for each i

and these are the functions we were looking for. o


As mentioned above, partitions of unity help to obtain global objects
by glueing local data. Here are some useful examples of this method.

Proposition 1.3. (1) (Bump functions) Let A c ~p be a closed set and


let U c ~p be an open neighborhood of A. Then there is a smooth function
() : ~p -+ [0, 1] that is identically 1 on A and identically 0 off U.

(2) (Uryshon separating functions) Let A, B c ~p be two disjoint closed


sets. Then there is a smooth function () : ~p -+ [0, 1] that is == 1 on A and
== 0 on B.

Proof. Assertions (1) and (2) are equivalent, taking U = ~n+l \ B. To


prove them, take V = ~n+l \ A and a smooth partition of unity {(), 'fJ} for
the covering {U, V}. Since () + 'fJ == 1 and 'fJ == 0 off V, it follows that () == 1
on A. Similarly, () == 0 off U, that is, on B. 0

Examples 1.4. We suggest that the reader represent in the one variable
case the function 'fJ in the proof of II.1.2, p. 50, and find the typical bump
or separating functions, whose graphs we depict below:
52 II. Manifolds

-€ o

In fact, these simple examples give a precise description of the set where
the value is == 1 or == 0 and can be used to modify other functions by
multiplication. For instance, if we multiply the function t t--+ t by the
separating function 0 in the right-hand figure above, or by 1 - 0, we get

Such modifications are used in many constructions. o

Another easy construction using partitions of unity is

Proposition 1.5 (C r Tietze Extension Theorem). Let A be a closed subset


of an open set U C JRP, and let f : A -+ JR be a cr function. Then f has a
Cr extension J: U -+ JR.

Proof. Choose local Cr extensions h : Ui -+ JR of flA n Ui so that the Ui


form an open covering of A in U. Add the open set V = U \ A and consider
a smooth partition of unity {1],Oi} for the cover {V, Ui}. Then J = L:i Oih
is well defined on the whole of U, because

(i) each term Oih vanishes on a neighborhood of U \ Ui, hence extends


by 0 off Ui, and

(ii) the sum is locally finite, as the family {Oi} is.


2. Differentiable manifolds 53

Furthermore, ! is Cr , because the local extensions h are, and the (}i'S


are smooth. Finally, for x E An Ui, (}i(x)h(x) = (}i(x)f(x), and the
same is true for x E A \ Ui also, since (}i(X) = O. Thus, for x E A,
/(x) = (L:i{}i(X»)f(x) = f(x). (Note that l1(X) = 0.) 0

Remark 1.6. Let X c jRP be an arbitrary set. The proof above can be
adapted to show that every C r mapping f : X -+ jRq has a C r extension to
some open neighborhood U of X in jRP. In other words, the notion of a C r
mapping, whose definition was of a local nature, is in fact global. 0

Exercises a nd problems

Number 1. Construct a differentiable function f : IR -+ IR such that f (t) = t for t $ 1,


f(t) ~ t for 1 $ t $ 2, and f(t) = 2 for t ~ 2.
Number 2. Exhibit two smooth functions f,9 : IR -+ IR that do not vanish on any
neighborhood of the origin but whose product does: f 9 == O.
Number 3. Exhibit two disjoint closed sets in the plane that cannot be separated by a
polynomial function.

Number 4. Let {£q be a smooth partition of unity for some covering U of an open set
U C IRP . Show that if all open sets of U have compact closure in U, then the smooth
function f = 2: i iBi : U -+ IR is proper.
Number 5. (Extension by zero) Let f : U -+ IR be a smooth function defined on a
neighborhood U of the origin in IRP . Show that there are smooth functions 9 : IRP -+ IR
that (i) coincide with f in some smaller neighborhood U' C U of the origin and (ii)
vanish off U.
Number 6. Let (ak) be a discrete sequence of pairwise distinct points in IRP and let (Tk)
be a sequence of integers ~ O. For every k and v = (VI, ... , vp) with Ivi = VI + ... + vp $
Tk, choose fkv E lR. Construct a smooth function f : IRP -+ IR with partial derivatives
a VIa at Vp (ak ) = fkv'
ivi
Xl ... Xp

2. Differentiable manifolds
Now that we have discussed differentiable mappings, we turn to the notion
of manifolds. We will only consider manifolds of finite dimension.

(2.1) Local diffeomorphisms. It is clear from the definition that a com-


position of Cr mappings is again a Cr mapping. But in general, the inverse
of a bijective Cr mapping need not be Cr again (think of t t---+ t 3 ). Thus we
54 II. Manifolds

say that a bijection f : X -+ Y is a Cr diffeomorphism when both f and


f- 1 are Cr mappings. The local version of this notion is essential: f is a
local Cr diffeomorphism at a E X when f is a Cr diffeomorphism from a
neighborhood of a onto another neighborhood of f(a).

The notion of local diffeomorphisms leads immediately to that of man-


ifolds.

(2.2) Manifolds. A subset M c jRP is a cr manifold when every point


x E M has an open neighborhood U in M which is Cr diffeomorphic to an
open set W in some Euclidean space jRm.
Here we fix some notation and terminology. Of course the terms dif-
ferentiable and smooth specify the class as they do for mappings. Now, a
given Cr diffeomorphism
tp = (tpl, ... , tpp) : W -+ U c jRP

is called a parametrization of M at x, or simply of U; the inverse diffeo-


morphism
x = (Xl, ... ,Xm ) : U -+ W C jRm
is called a local coordinate system. Given a second parametrization 'I/J, the
composite mapping 'I/J-l 0 tp is the change of coordinates, which is a Cr
diffeomorphism of open sets in Euclidean spaces. Hence its derivative is
a linear isomorphism, and the dimension m of the domain of local coordi-
nates is the same for the two parametrizations: this m is the dimension of
M at x: m = dimx(M). If the dimension is the same at all points, we just
write m = dim(M); unless otherwise stated, we always assume this. Also,
as customary, we will say that M c jRP is a curve when m = 1, a surface
when m = 2, and a hypersurface when p = m + 1.

Thus, manifolds are sets locally diffeomorphic to Euclidean spaces and,


in particular, locally homeomorphic to them. Consequently, manifolds
share with Euclidean spaces all topological properties of a local nature:
local connectedness, openness of connected components, arc-connectedness
if connected, local compactness,.... Of course, since our manifolds are by
definition embedded in Euclidean spaces, they inherit many other proper-
ties: metrizability, countable bases of open sets. In particular, manifolds
are paracompact spaces. Moreover, combining everything, manifolds have
other properties, such as exhaustion by compact sets.

(2.3) Equations. A basic property of a differentiable cr manifold Me jRP


2. Differentiable manifolds 55

is that it is locally closed (open in its closure) in ~p. In fact, there are local
cr equations at every point x E AI: there are Cr functions JI, ... , fq defined
on an open neighborhood U of x in ~p, such that:

(i) Un M = f-l(O) = {x E U: JI(x) = ... = fq(x) = O}.


(ii) The rank of the Jacobian matrix (;~i.) is q at every point x E UnM.
J
(iii) dimx(M) = p - q.

In fact, starting from a parametrization at x, one constructs, using the


Inverse Mapping Theorem, a Cr diffeomorphism h = (hI, ... , hp ) from U
onto an open set V C ~p such that h(U n M) = V n (~m X {O}). Then
the equations above are Ii = h m +i . The existence of h simply says that the
pair M c ~p is locally cr diffeomorphic to the pair ~m C ~p.
Conversely, if a subset M C ~p can be locally described by functions
verifying (i) and (ii) above, then it is a manifold and its dimension is given
by (iii).
It is important to stress that a manifold M need not have global equa-
tions, that is, functions as above defined on U = ~p; indeed, condition (ii)
is highly restrictive. We will discuss this matter for hypersurfaces, as it
is narrowly related to the Jordan Separation Theorem (II.7.5, p. 88, and
III.6.4, p. 129).

Example 2.4. The first examples of manifolds are spheres:

§m = {x E ~m+1: LX; = I}.


i

They can be parametrized by the so-called stereographic projections; in fact,


two such projections suffice. Let aN = (0, ... ,0,1) E ~m+1 be the north
pole of §m, and as = (0, ... ,0, -1) the south pole. Then the projections
from the north and south poles,

7rN:§m\{aN}-+~m:Xt--+(I Xl ""'1 Xm )
- Xm+l - Xm+l

and
7rs : §m \ {as} -+ ~m : x t--+ ( Xl , ... , Xm ) ,
I+Xm+l I+xm+1
are local coordinate systems, corresponding to the parametrizations
56 II. Manifolds

and

Y= (YI,"" Ym) H
( 1 +2YIIIyI1 2, ... , 1 +2Ym 1-IIYI12)
IIyI1 2' 1 + IIyI12 .
On the other hand, spheres do have global equations: the expression
Ei x~ = 1 is a global equation in the sense of II.2.3, p. 55. 0

Let us point out here that the abstract approach where M is not a subset
of any jRP is not so abstract: just assuming its topology has a countable
basis, M embeds in some jRP. This can be done by elementary means, if one
disregards how big p is; we do not appeal here to the Whitney Embedding
Theorems that control p to 2 dim(M) + 1.

(2.5) Manifolds with boundary. The definition of manifold stems from


the choice of a local model: Euclidean spaces. The choice of this model
is convenient because Euclidean spaces are the natural setting for differ-
ential analysis. However one can go a little further and extend without
surprises differential analysis to Euclidean half-spaces, namely to convex
sets lHIm c jRm defined by one linear inequality ,X{x) ~ O. After linear
changes of coordinates, we get the canonical lHIm described by Xl ~ 0 in
coordinates X = (Xl, ... ,Xm ) E jRm. Then, a cr manifold with boundary
is a set M C jRP locally Cr diffeomorphic to such half-spaces. The notions
of parametrization, local coordinate system, and dimension are the same.
Furthermore, since the half-spaces ,X ~ 0 have quite obviously the boundary
,X = 0, the more general new manifold M has its own boundary aM, which
consists of the points X at which 'x{x) = 0 in some coordinate system. The
points in M \ aM are usually called interior points of M, but this has no
relation with the interior of M in jRP.
In the end, all of this is consistent because a diffeomorphism of half-
spaces preserves their boundaries. We remark here that this is true by the
differentiable version of the Invariance of Domain Theorem. This differen-
tiable version admits a thoroughly elementary proof, in dramatic contrast
with the general theorem, which we will prove later twice (IV.5.6, p. 167,
and IV.7.5, p. 181), by means of Euclidean degree theory.
Thus we have boundaryless manifolds and manifolds with boundary. In
particular, if M is a manifold with boundary, of dimension m, then aM
is a boundaryless manifold of dimension m - 1, and M \ aM is also a
boundaryless manifold of dimension m.
Finally, we remark that the arguments of II.2.3, p. 55, applied to one
single differentiable function f : U ---+ jR, show that if the partial derivatives
2. Differentiable manifolds 57

3h do not vanish simultaneously at any x E U with f(x) = 0, then the


inequality f 2 0 defines a manifold with boundary defined by f = o.

(2.6) Product of manifolds. It is clear that the product of two manifolds


is also a manifold, at least if boundaries do not interfere. In fact, we will
only consider (i) the product of boundaryless manifolds, again boundary-
less, and (ii) the product of a manifold N with boundary and a boundaryless
manifold M, which has the boundary a(N x M) = aN x M. For instance,
if N = [0,1] is an interval, we get the disjoint union of two copies of M,
namely

a([0, 1] x M) = Mo U MI, Mo = {O} x M, Ml = {1} x M.

It is also clear that the dimension of a product is the sum of the dimensions
of the factors, and the class of the product is the smallest class of a factor.

(2.7) Smoothness. The study of mappings of cr manifolds is not re-


stricted to those of class cr. In fact, the main interest concerns continuous
mappings, not even differentiable. However, dealing with the class Cr in-
volves quite a bit of difficulty, mainly due to the loss of differentiability that
occurs in many constructions. This problem disappears completely for class
00, and we will mainly take the approach of restricting our presentation to
smooth manifolds.
This restriction is supported by the fact that actually all manifolds are
smooth. This means that by looking only at smooth manifolds, we just de-
prive the reader of some technical diversion, but all results we prove apply
to arbitrary manifolds. In any case, the consideration of finite class map-
pings is of relevance even on smooth manifolds and is actually a crucial
matter for extending the theory to arbitrary cr manifolds. We will confine
this discussion to a special section in the proper place (111.4), always in a
concentrated form to remain faithful to our general continence policy con-
cerning technicalities.

The essential concept that makes manifolds the suitable object to de-
velop calculus is that of tangency.

(2.8) Tangent space. Let M be a differentiable manifold, and let x E M.


Pick any parametrization <p : W -+ M at x (with W c ]Rm or lHIm ), and let
a E W be such that <p(a) = x. This <p is in fact a differentiable mapping
W -+ ]RP, and we have its derivative da<p : ]Rm -+ ]RP. One easily checks
that the linear space im(da<p) C ]Rm does not depend on the choice of <p:
58 II. Manifolds

it is the tangent space to M at x and is denoted by TxM. The vectors in


TxM are called tangent to M at x.
It is easy to produce a basis for TxM. Just consider the canonical basis
ei = (0, ... ,1, ... ,0) of]Rm and set

Notice that this basis ofTxM consists of the columns ofthe Jacobian matrix
of da'P : ]Rm -+ ]RP. The partial derivative notion comes from the fact that
one can look at any vector u E ]Rn as a directional derivative. We will
not pursue this fruitful viewpoint here, but we keep the notation of partial
derivatives anyway.
In a different way, one can use local equations to describe the tangent
space. Namely, if II, ... ,fq are local equations of M in a neighborhood U
of x (11.2.3, p. 55), then u E TxM if and only if

(this is generalized later in II.3.2, p. 62). For one single equation f = II,
that is, when M is a hypersurface in ]Rm+1, we have

af af ~
0= dxf(u) = (gradx(f), u), gradx(f) = ( aX} (x), ... , axp (x);;

that is, TxM is the linear hyperplane perpendicular to the gradient grad x(f).
Finally, note that for a product M x N of two manifolds we have

Example 2.9. Let us look again at spheres §m. One can use the stereo-
graphic projection 7rN to obtain a basis Tx§m at the south pole x = as:

Similarly, 7rS gives the same basis of Tx§m at the north pole x = aN. We
see that Tx§m ..L x in both cases. This is in fact true for all x E §m, and we
invite the reader to pursue the computations at an arbitrary x. However,
it is much easier to use the equation f(x) = I:i xT - 1 = 0: TxM is the
hyperplane perpendicular to the gradient gradx(f) = 2x. 0
2. Differentiable manifolds 59

Now we can define the following:

(2.10) Derivative of a differentiable mapping. Let f : M ~ N be


a differentiable mapping of differentiable manifolds, and consider a point
x E M. We can extend f to a differentiable mapping f from some open
neighborhood of x in RP into some open neighborhood of y = f(x) in Rq,
and this extension has a derivative dx ! : RP ~ Rq. Then, it is easy to check
that by restriction we get a well-defined linear mapping

which does not depend on the choice of f and is called the derivative of f
at x.
Alternatively, we can localize, that is, pick coordinate domains U at
x and V at y such that f (U) c V, and build the following commutative
diagrams:

U ~ V TxM ~ TyN
da'P r~ ~ 1
(db'I/J)-l

Rm ~ Rn
(where <p is the parametrization of M at x, with <p(a) = x, and '!jJ is that
of N at y, with '!jJ(b) = y). One sees that the matrix of the linear mapping
daf with respect to the bases of the partial derivatives in both spaces TxM
and TyN is the Jacobian matrix of

We insist that the consideration of manifolds is necessary to have deriva-


tives in any reasonable sense: the notion of differentiable mappings was
given from the very beginning with no restriction on the sets involved, but
for derivatives we need tangent spaces, available only for manifolds. Then,
once derivatives are defined, calculus follows: chain rule, etc. The only
caution concerns the boundary sometimes. For instance, it is easy to see
that if a differentiable function f : M ~ R has a local extreme at a point
x E M \ aM, then the derivative dxf : TxM ~ R is identically zero.

In the same vein, the Inverse Mapping Theorem must be modified to


read: a differentiable mapping f : M ~ N is a local diffeomorphism at x if
and only if dx f is a linear isomorphism and f preserves boundaries locally
at x. Of course, if M is boundaryless, the latter condition is meaningless.
60 II. Manifolds

(2.11) Tangent vectors and curve germs. We can describe tangencies


in a very geometrical way: a tangent vector to M at x is always tangent
to some curve germ in M. This means that for every u E TxM there are
differentiable mappings, : [0, c) -+ M c jRP such that ,(0) = x and

u = ,'(0) = (ri (0), ... , ,~(o» = do,(l) E TxM c jRP.

Furthermore, if f : M -+ N is differentiable, dxf(u) = (f 0 ,)'(0).


Example 2.12. Here there is a simple application of tangent vectors: if a
curve and a manifold meet transversally at a point, then the point is isolated
in their intersection.
Let r be a curve and M a manifold, both in jRP, and let a Ern M.
The hypothesis says that no tangent vector to r at a is tangent to M at a.
For the proof we can suppose that a is the origin in M = jRm X {o}. Then
choose a parametrization, : t H (ri(t» of r with ,(0) = a, and we know
that ,'(0) ~ TaM = jRm X {o}. Hence ,~(o) =1= 0 for some i > m, so that
,i(t) is strictly monotonous near 0 and 0 is an isolated zero. This means
that ,(t) ~ M = jRm X {o} for small enough t =1= 0, as claimed. D

Exercises and problems


Number 1. Prove that a subset M C ]RP is a differentiable manifold if and only if for
every point x E M there are an open set A in some half-space IHlm and a differentiable
mapping t/J = (t/Jl,'" ,t/Jp) : A -+ ]RP such that (i) t/J is a homeomorphism onto some open
neighborhood of x in M and (ii) the Jacobian matrix (~) has rank m at t/J-l(X).
J

Number 2. Show the following:


(1) The cusp in ]R2 given by the equation x 2 = y3 is not a smooth curve, but it is
homeomorphic to ]R.
(2) The semi-cone in ]R3 defined by x 2 + y2 = Z2, Z ~ 0, is not a smooth surface, but
it is homeomorphic to ]R2.

Number 3. (1) Show that the solid torus M C ]R3 generated by the disk y = 0,
(x - 2)2 + z2 ::; 1 around the z-axis is a manifold with boundary the torus generated by
the circle y = 0, (x - 2)2 + Z2 = 1.
(2) Show that the subsets Ml : xi + x~ ::; ~ and M2 : x~ + x~ ::; ~ of the unit sphere
§3 C ]R4 : xi + x~ + x~ + x~ = 1 are diffeomorphic to M.
(3) Deduce that §3 is the union of two solid tori along their boundaries, so that the
meridians of each one are the parallels of the other.

Number 4. Let M be a Hausdorff topological space M with a countable basis of open


sets. A chart on M is a homeomorphism x : U f-t ]Rm from an open set U of M onto
an open set x(U) C ]Rm. Two charts x and y with domains U and V are called cr
2. Differentiable manifolds 61

compatible (we fix the differentiability class r henceforth) when the mapping (of open
sets in Euclidean spaces)

yo x-I: x(U n V) ---t y(U n V)


is differentiable. A differentiable atlas in M is a collection of Cr compatible charts whose
domains cover M; when such an atlas exists, we say that M is an abstract (boundaryless)
manifold.
(1) Check that the concrete manifolds M C ]RP defined in the text are also abstract
manifolds. Check that dimension is defined well and is the same in the abstract as in the
concrete setting.
(2) Show that real and complex projective spaces lRlP'm and cpm are abstract smooth
manifolds, using homogeneous coordinates (xo : ... : x m ) to define charts with domains
Xi i= O. Compute dimensions.

Number 5. A continuous mapping f : M ---t N of abstract manifolds is Cr if for every


two charts x, y on M, N, with domains U, V such that f(U) C V, the localization yo
f 0 X-I: X(U) ---t y(V) is cr. Then we have the notion of diffeomorphism, and the loop
closes, because every abstract manifold M is diffeomorphic to some concrete manifold.
(1) Check that differentiable mappings between concrete manifolds as defined in this
section are exactly differentiable mappings in the abstract sense.
(2) Prove that ]Rpm is diffeomorphic to a compact manifold in some ]RP, using the
mapping

How can this be adapted to cpm?


(3) Show that the real projective line is diffeomorphic to §l and the complex projec-
tive line is diffeomorphic to §2.
Number 6. Let p : ]Rn+l \ {O} ---t §n be the radial retraction given by x t-+ xlllxll. For
every point a E ]Rn+l, a i= 0, let Sa be the sphere centered at the origin with radius
lIall, and denote Ha = TaSa. Show, without explicit computations, that daP(a) = 0
and daPIHa is the homothecy of ratio 1/11all. Deduce from this an expression for the
derivative daP(u).
Number 7. Let k 2: 1 be an integer. Show that the set Mk C ]R3 given by X 2k +
y2k + Z2k = 1 is a smooth surface and that the radial retraction X t-+ xlllxll induces a
diffeomorphism from Mk onto the unit sphere §2.
Number 8. Let M C ]RP be a differentiable Cr (r 2: 2) manifold of dimension m. Prove
that the tangent bundle

TM = {(x,u) E M x]RP : U E TxM} C]RP x]RP

is a differentiable cr- 1 manifold of dimension 2m. Show also that the Stiefel bundle

VM = {(X,Ul,'" ,Urn) EM x]RP x··· x]RP: Ul, ... ,Urn is a basis of TxM}

is a non-compact differentiable Cr - I manifold of dimension m(m + 1).


62 II. Manifolds

3. Regular values
In this section we come to the key notion behind most constructions in
differential topology. We refer to transversality. However, we do not need
it in full generality, and for this reason we restrict our discussion to regular
values, which is the first instance of transversality.

(3.1) Regular values. Let f : M -t N be a differentiable mapping. A


critical point of f is a point x E M whose derivative dxf : TxM -t Tf(x)N
is not surjective. Critical points form a closed set, because the condition on
their derivatives can be expressed by some equations involving the Jacobian
determinant of a localization of f. The non-critical points are called regular
points, and they form an open set.
Now we look at a E N: we say that a is a critical value of f if there is
some critical point x E M with f(x) = a. Hence, the set of critical values
of f is the image of that of its critical points but need not be closed (unless
f itself is closed, e.g., when M is compact). If a E N is not a critical value,
then we call it a regular value of f.
We denote by Cf C M the set of critical points of f and by Rf C N
the set of regular values of f. Hence Rf = N \ f(Cf).
Notice that if dim(M) < dim(N), then Cf = M and Rf =N \ f(M).

Now we recognize that when we defined equations II = ... = fq = 0 of


a manifold Me ]RP on an open set U C ]RP (11.2.3, p. 55), we just said that
owas a regular value of the mapping f = (II, ... , fq) : U -t ]Rq. In fact, the
main feature of regular values is that their inverse images are manifolds.
This is an easy consequence of the Inverse Mapping Theorem but must be
stated with some care to keep track of boundaries:

Theorem 3.2. Let f : M -t N be a differentiable Cr mapping, and let


a E N \ 8N be a regular value of both f and flaM. Then
(1) f-l(a) is a manifold with boundary f-l(a) n 8M, whose dimension
is dim(M) - dim(N).
(2) Txf-l(a) = ker(dxf) for every x E f-l(a).

Concerning the proof of this result we will at least mention the following.
For every x E f-l(a) there are coordinate systems (Xl, ... , xm) of M at x
and (Yl, ... , Yn) of N at a such that Yi 0 f = Xm-n+i for i = 1, ... , n (and
3. Regular values 63

Xl 2:: 0 where the boundary is involved). In other words, the localization of


f near x is a linear projection (Xl, ... , Xm) H (X m- n+1,"" xm). Then

{
dxf(a X m a- n + .1)x = 1..1
1 VI'" a
for i = 1, ... , n, and
a~j Ix' j = 1, ... , m - n, generate Txf-l(a).

In fact, (Xl, ... , xm - n) is a coordinate system of our inverse image at x.

We will use this result mainly when dim(M) - dim(N) = 1, so that


f-l(a) is a curve. Then it is important to know that curves always have
global parametrizations:

Theorem 3.3 (Classification of compact differentiable curves). Let C be


a compact Cl curve. Then C is Cl diffeomorphic to either the unit circle
§l C JR2 or the closed interval [0,1] c JR.

Of course, to use these theorems we need to find regular values, but this
is the famous Sard-Brown Theorem:

Theorem 3.4 (Sard-Brown Theorem). Let f : M -+ N be a differentiable


Cr mapping, with r > dim(M) - dim(N). Then the set of regular values of
f is residual (a countable intersection of dense open sets), hence dense, in
N.
In particular, if dim(M) < dim(N), then N \ f(M) is residual, hence
dense, in N.

The last assertion of the theorem is sometimes called the Easy Sard
Theorem, because it can be proved directly by an easier argument.
We remark that the proof of this theorem for finite class r < +00 is
technically very demanding, as it requires the K neser- Glaeser Theorem to
restore the differentiability class lost by partial derivation. Of course, this is
no problem in the smooth case r = +00, and the proof is much easier in this
case. Furthermore, this is enough for us, as we will reduce the differentiable
case to the smooth one. In fact, as mentioned before (11.2.7, p. 57), every
differentiable manifold has a smooth model, and this deep result roughly
says that our limitation to smooth manifolds is only technical.
After all these considerations, we include here the following proof:
Proof of the Sard-Brown Theorem for smooth mappings. First of all notice
that since M c JRP is locally compact, we can apply the Baire Theorem,
64 II. Manifolds

so that any countable intersection of dense open sets of M is again dense,


which is the last remark in the statement.
Since manifolds have countable bases of open sets, one easily sees that
the statement is of a local nature; hence we can assume that M, N are
open sets of Euclidean spaces or half-spaces and that f is the restriction
of a smooth mapping g : U -+ IRq defined on an open subset U of IRP.
Clearly, the result for f follows from the result for g; hence we can simply
suppose f = g (this means we can disregard boundaries). We will prove
the following statement, which in fact belongs to measure theory:

(*) The image of the set C, of critical points of f has measure zero.

Recall that a set has measure zero when it is contained in a countable


union of cubes whose total volume can be taken arbitrarily small. It is easy
to see that a countable union of measure zero sets also has measure zero,
so that being measure zero is a local question. It is also immediate that a
measure zero set has empty interior. In our case, this implies that R, is
residual. Indeed, C, is closed, hence a countable union of compact sets, so
that f(C,) = IRq \ R, is also a countable union of compact (hence closed)
sets, each one contained in f(C,), hence each one with empty interior.
We now prove the above assertion (*). The proof is Milnor's by in-
duction on p. Assume the result for smooth mappings with domain in
Euclidean spaces of dimension < p. Denote

Ci = {x E U: all partial derivatives of order :S i vanish at x}

and consider the sequence of closed sets

We split the argument into several steps.

(a) The set f(C, \ Cd has measure zero.

We will see that every point a E c, \


Cl has a neighborhood V such that
f(C, n V) has measure zero. As we can cover C, \ C l by count ably many
such neighborhoods, this implies (a).
Fix a and choose some partial derivative of f that does not vanish at a,
say ~(a) i= o. Then h(x) = (h (x), X2, . •. , xp ) is a local diffeomorphism
at a, and we can replace f by f 0 h- l on a neighborhood V of a. In
other words, we can suppose h(x) = Xl on V, so that f preserves every
3. Regular values 65

hyperplane Xl = t; we denote by 9 the restriction of f to that hyperplane.


Notice that for X = (t, x') E V it is

Df(X)=(! 0 )
Dg(X')

so that if D f(x) is not surjective, Dg(x' ) is not either. This means that if
x E Cf n V, then x' E Cg • Thus we have

f(Cf n V) n {Xl = t} C f(Cf n {Xl = t}) C g(Cg ),

and the latter set has measure zero by induction. Summing, f(Cf n V)
cuts each hyperplane Xl = t on a measure zero set, and the Fubini Theorem
(which has an elementary proof for measure zero sets) guarantees that
f(Cf n V) has measure zero, as desired.

(b) Each set f (Ci \ Ci +I) has measure zero.

As before, we see that every point a E Ci \ CHI has a neighborhood V such


that f (Ci n V) has measure zero. By definition, there is a partial derivative
w of f of order i that vanishes at a and such that 8w(a)/8xj =1= 0; for
ease of notation, we suppose j = 1. Then h(x) = (w(x), X2, ... , xp) is a
local diffeomorphism at a, and on some neighborhood V of a the inverse
mapping h- l : W -+ V is well defined and smooth. We have h(Ci n V) c
{Xl = O}, and every point in h(Ci n V) is a critical point of the restriction
9 = f 0 h-IIW n {Xl = O}. Hence, f(Ci n V) = g(h(Ci n V)) c g(Cg ), and
the latter set has measure zero by induction. We have proved (b).

(c) Ifi > p/q -1, the set f(Ci) has measure zero.

Consider a compact cube K C U of side 8 > 0, and let us see that f (Ci n K)
has measure zero. Since Ci can be covered by countably many such K's,
the conclusion follows. Now, by Taylor expansion and the definition of Ci,
we have
f(y) = f(x) + R(x, y), IIR(x, y)1I ~ clly - xW+ I
for any x, y E Ci n K. The constant c depends only on f and K. Now,
subdivide K into n P cubes L with side 8/n and diagonal JP8/n. Choose
X E C i n L, so that for every y E L,

. ~ P = c(JP8)HI
Ilf(y) - f(x)1I ~ IIR(x, y)11 ~ clly - xll~+l ----:;;- .
66 II. Manifolds

Thus, f(L) is contained in the cube L* with center f(x) and side 2p. Con-
sequently, f(Ci n K) is contained in the union of n P cubes with side 2p,
whose total volume is
c'nP- q (i+1) .

Since i > p/q - 1, this volume goes to zero when n -+ 00. This proves (c).
It is clear that the three facts (a), (b), and (c) imply (*), and this com-
pletes the proof of the Sard-Brown Theorem for smooth mappings. 0

Exercises and problems

Number 1. Let P be a homogeneous polynomial of degree d in n + 1 indeterminates.


(1) Prove that
n+1 oP
LXiOX = d·P.
i=l 1.

(2) Show that every equation P = e:, e: t= 0, defines a smooth hypersurface Me in


jRn+l (except if there are no zeros).
(3) Prove that Me is diffeomorphk to Ml for e: > 0 and to M-l for e: < O.
(4) Also, must Ml and M-l be diffeomorphic? How does d affect this?
(5) What about the equation P = O?
Number 2. Let M C jRn be a non-empty real quadric given by

L aijXiXj = 1,
l$:i,i$n

with det(aij) t= o. Show that M is a smooth hypersurface diffeomorphic to a product


§k x jRt, for suitable k and f.

Number 3. Prove that the set M C jRrnxn of all matrices of rank k is a smooth manifold
of codimension (m - k)(n - k).
Number 4. Let E C jRnxn be the linear space of all symmetric matrices of order n, and
consider the smooth mapping f : jRnxn -+ E : A I-t At A.
(1) Show that dAf(B) = At B + Bt A, A, B E jRnxn, and deduce that the orthogonal
group O(n) = f-l(I) C jRnxn is a compact manifold of dimension n(n - 1)/2.
(2) Prove that O(n) has two diffeomorphic connected components, one, SO(n), de-
fined by the condition det = +1.
(3) Compute the tangent space to O(n) at A = I. At which other A E O(n) is the
tangent space the same?
Number 5. Let M C jRP be a differentiable cr (r ::::: 2) manifold of dimension m. Prove
that the (orthonormal) Stiefel bundle

OM = {(x, Ul, ... ,Urn) E M xjRP x· .. xjRP : Ul, ... ,Urn is an orthonormal basis of TxM}

is a differentiable Cr - 1 manifold of dimension ~m(m + 1), compact if and only if Mis.


(Compare with Problem 8 in 11.2.)
4. Thbular neighborhoods 67

Number 6. Let U C IRm be an open set, and let f : U -+ IRn be a smooth mapping,
m ~ n. Suppose that 0 is a critical point and f(O) = O. Consider the matrices m x n as
elements of IRmxn and consider

F: (U \ {O}) x IRmxn -+ IRn : + Ax.


(x, A) f-t f(x)
(1) Check that the origin is a regular value of F, hence that M = F- 1 (0) is a smooth
manifold.
(2) Show that if A E IRfflxn is a regular value of the projection 7r : M -+ IRmxn , then
the mapping
FA : U -+ IRn : x f-t f(x) + Ax
has no critical point x =1= 0 with value FA(X) = O.
(3) Conclude that there are A's with entries arbitrarily small such that the origin is
the only critical point with value o.

Number 7. A real projective hypersurface of degree k is a subset Me lU'm defined by


a homogeneous equation P(xo, .. . ,xm ) = 0 of degree k.
(1) Under what conditions on P is M a smooth manifold?
(2) Show that if k = 2£ is even,

f m ( ) P(xo, ... ,xm )


: lU' -+ IR: Xo:···: Xm
Xo + ... +xm2 )l
f-t ( 2

is a well-defined smooth mapping. When is 0 a regular value of f?


(3) What about real projective hyperplanes H C lU'm, which are defined by linear
homogeneous equations?

4. Tubular neighborhoods
Here we recall the constructions of tubular neighborhoods. These are a
key ingredient in the next section for the approximation results that reduce
homotopy to smooth homotopy, but, also, we will need them in a crucial
way to prove the Poincare-Hopf Index Theorem at the very end this book
(V.7.2, p. 219).
The essential fact is that every differentiable boundaryless manifold in
a Euclidean space ]RP is the differentiable retraction of some neighborhood
of it in ]RP. We first take a purely topological approach that preserves
differentiability class well:
Proposition 4.1. Let M c ]RP be a boundaryless differentiable Cr mani-
fold. Then there is an open neighborhood U of M in]RP and a Cr retraction
p: U -+ M (that is, a Cr mapping with p(x) = x for x EM).

Proof. Since the pair M c ]RP is locally Cr diffeomorphic to the pair ]Rm c
]RP (II.2.3, p. 55) and linear projections ]RP -+ ]Rm are smooth retractions,
each point x E M has an open neighborhood U in ]RP equipped with a
68 II. Manifolds

cr retraction p : U -+ U n M. Henceforth the proof consists of a careful


glueing of such local retractions. We split the argument into several steps.
First we note that since M is locally closed in jRm, it is closed in some open
neighborhood n in jRP; for the rest of the proof, all closures are considered
inside this open set n. Also, all mappings are cr.
Step I: Clueing two local retractions. Let a : V -+ V n M and r : W -+
W n M be two retractions defined on open sets V, W of n. Then there is
a retraction
p : U -+ un M = (V u W) n M.

If T = V n W n M = 0, there are disjoint open sets VI, WI c n such


that V n M C VI C V and W n M C WI C W. Indeed, in the metric
space V U W, the two disjoint closed sets V \ w, W \ V can be separated
by disjoint open sets VI C V, WI C W. Now, we take U = VI U WI, and p
is the trivial glueing of two mappings defined on disjoint open sets. After
this remark we assume T =1= 0.
The idea is to glue a and r through a homotopy that deforms one
retraction into the other over T. To that end, consider D = a-I(T) n
r-I(T). This set is open in n, hence in jRP, and the set

E = {x ED: D:J [a(x), r(x)]} :J T

is open too. Indeed, suppose D contains a segment [a(x),r(x)]. Since D


is open in jRP, it contains a convex open neighborhood Q of that segment,
and then a-I(Q) n r-I(Q) C E is an open neighborhood of x. Then we
can define mappings H t , 0 :S t :S 1, on E by

Next, notice that the disjoint sets V n M \ Wand W n M \ V are closed in


V U W. For instance V n M is closed in V, and hence V n M \ W is closed
in V \ W, which is closed in V U W. Consequently, we can find a smooth
bump function 8, which is == 0 on a neighborhood Uu C V of V n M \ W
and == 1 on a neighborhood UT C W of W n M \ V. Finally, we define the
retraction p on the open set U = Uu U E U UT by

a(x) on Uu ,
p(x) = { H(J(x) (x) on E,
r(x) on UT •

Some straightforward computations show this construction is consistent.


4. Tubular neighborhoods 69

Step II: Exhaustion of M by local retractions. There is a sequence of re-


tractions Pk : Uk --+ Mk = Uk n M such that

M = UMk and Mk C Mk+l·


k

As mentioned before, we can cover M by local retractions T : V --+


V n M, and then, since M has a countable basis of open sets, we get a
sequence Tk : W k --+ Wk n M, with M c Uk Wk. Now, by Step I, for each
k we can glue Tk and Tk-l, and Tk-2, and so on, to obtain a retraction
Uk : Vk --+ Vk n M, with

Next, set
Mk = {x EM: dist(x,M\ Vk) > i}.
This is an open subset of Vk n M, and

because Vk C Vk+l. On the other hand, for every x E M, say x E Wk C Vk,


we have

dist(X, M \ Vi) ~ dist(x, M \ Vk) for.e ~ k (as Vi n M ~ Vk n M),


{
dist(x, M \ Vk) > i for .e large (as dist(x, M \ Vk) > 0),

and thus x E Me for large.e. Consequently, M = Uk Mk. Finally, take


Uk = u;l(Mk) and Pk = ukluk.
Step III: Splitting into two sequences of disjoint retractions and conclusion.
First, we shrink the domains of the retractions Pk of Step II as follows.
Take targets

MlI = Ml, M2I = M2, MkI = Mk \M


-
k-2 for k ~3

and domains U~ = p;l(M~). The virtue of this is that the M~'s still cover
M, since M k C Mk+l, but besides, we have

M~ n Mi = 0 if k ==.e mod 2
(and k =f: .e, of course). Indeed, suppose .e = k + 2n. Then k ~ .e - 2, and

{ M~ C Mk c Me-2'
Mi = Me \ Me-2,
70 II. Manifolds

hence our assertion. Now put

V' = UU~e-l' W' = UU~e'


e e

These two sets are open, and M c V' U W'. Let us look at V' first. The
sets M~e-l = U~e-l n M c V' n M are closed in V' and form a discrete
family.
Indeed, suppose that some x E M~eo-l is the limit of a sequence of
points Xe E M~e-l c M. Then Xe E U~eo-l for e large enough; hence
Xe E U~eo-l n M = M~eo-l' a contradiction.

Consequently, the sets M~e-l can be separated by disjoint open sets


A 2e- 1 C V', and P2e-l restricts to a retraction U~e-l n A 2e- 1 -+ M~e-l'
All these restrictions glue trivially, because their domains are disjoint, to
give another retraction

e e

Similarly, starting with W', we get a second retraction

Peven : U(U~e n B 2e ) -+ UM~e'


e e

and we are done, because by Step I we can glue these two retractions to
obtain P : U -+ M, the one we sought. D

The preceding proof shows that being a retract of an open set in ~p is a


local property. Hence, the differentiable content of the last theorem is the
basic fact that manifolds are locally retracts. Actually, this local property
characterizes manifolds:

Proposition 4.2. Let M c ~p be a subset such that every point x E M


has an open neighborhood U in ~p on which there is a differentiable cr
retraction P : U -+ U n M. Then M is a boundary less cr manifold.

Proof. Fix x E M and a Cr retraction P : U -+ un M, with x = p(x) E


unM. After a translation, we can assume x is the origin. Denote 7r = dop,
and define
h = (ld]Rp -7r) 0 (ld]Rp -p) + 7r 0 p.
Since p = pop, we have

7r = dop = do(p 0 p) = 7r 0 7r,


4. Tubular neighborhoods 71

which readily gives doh = IdlR.p. Hence h is a Cr diffeomorphism on an


open neighborhood V C U of the origin. Further computation shows that
hop = 7f 0 h; hence

h(W n M) = h(p(W)) = 7f(h(W))


for W = V n p-l(V). But 7f(h(W)) is open in L = 7f(~P), because linear
maps are open. Thus W n M is Cr diffeomorphic to an open set of a linear
space; that is, W n M is a Cr manifold. D

There is a second and more popular method to construct retractions.


We present this method now for smooth manifolds. As usual, we denote by
-L perpendicularity in Euclidean spaces.

Proposition and Definition 4.3. Let M C ~p be a boundaryless, smooth


manifold of dimension m. Then the set

vM = {(x,u) E ~p x ~p: u -L TxM, x E M}

is a boundaryless, smooth manifold of dimension p, and the smooth mapping

{} : vM --+ ~p : (x, u) H x + u.
induces a diffeomorphism from an open set fl :J M x {O} onto another
U :J M. Consequently, we have the smooth retraction

p: U --+ M : y = {}(x, u) H X

and it follows that (maybe for a smaller U):


(1) M is closed in U,
(2) y - p(y) -L Tp(y)M for all y E U, and
(3) dist(y, M) = Ily - p(y) II for all y E U.
In fact, (3) characterizes p.
We summarize the situation by saying that p : U --+ M is a tubular
retraction of M in ~p.

Proof. To start with, we prove that


(a) vM is a boundaryless, smooth manifold of dimension p.
Consider any local equations for M, that is, differentiable functions
II, ... , fq, q = p - m, defined on an open set V C ~P, such that:
72 II. Manifolds

(i) VnM={xEV: JI(x)= .. ·=fq(x)=O} and


(ii) the rank of the Jacobian matrix (;~i.) is q at every point x E VnM.
J

Condition (i) means that the gradients

are perpendicular to TxM, and condition (ii) means that they are indepen-
dent. Hence, they generate TxM.1. Consequently we can define a bijection

(V n M) x IRq -+ (V x JRP) n vM: (x,y) M (x,u)

by
q

u = LYi gradx(Ji).
i=l

This is in fact a diffeomorphism: its inverse is obtained by solving the linear


system (*) in the unknowns Yi. Since we can cover M by open sets like V,
we see that v M is locally diffeomorphic to M x IRq, which is a manifold of
dimension m + q = p.
Next we see that
(b) iJ is a local diffeomorphism at every point (x, 0).
Indeed, it is enough to check that d(x,o)iJ is a linear isomorphism. But
iJ induces by restriction the identifications

iJIMx{O} : M x {O} == M and iJl{x}xT",Ml.: {x} x TxM.1 == TxM.1,

so that the image of d(x,o)iJ contains TxM + TxM.1 = IRP. Hence, that
derivative is surjective, and since dim(vM) = dim(IRP), we are done.
Once we know that iJ is a local diffeomorphism around M x {O}, what
remains to show is that
(c) iJ is injective on some neighborhood of M x {O}.
Clearly, iJ is injective on the set

E = {(x, u) E vM: Ilull < T(X)},


where T is defined as follows:
4. Tubular neighborhoods 73

T(X) is the infimum of all s E JR. such that there are (x, u), (y, v) E
• vM with Ilvll :s; Ilull = s, x =1= y, 19(x, u) = 19(y, v).
Thus, we must check that E is a neighborhood of M x {o}.
Fix z E M. We already know that 19 is injective in a neighborhood V
of (z, 0), say V : IIx - zll + lIull < E for a suitable E > O. We claim that
T(X) ~ 1E if Ilx - zll < ~E.
For, suppose IIx-zll < ~E and let (x, u), (y, v) in the definition of T(X) verify
1
Ilvll :s; Ilull < E• Then (x, u) E V and (y, v) ¢: V; moreover, IIx - yll =
Ilu - vii :s; 211ull· Consequently:
iE > 211ull ~ Ilx - yll ~ Ily - zll - Ilx - zll
~ (E -lIvll) - ~E > (E -1E) - ~E = ie,
a contradiction.
Now, the claim implies
E:J {(z,u) E vM: liz - xii < ~e, lIull < 1e},
and the latter set is an open neighborhood of (x, 0).
Consequently, there is some open neighborhood n of M x {O} on which
19 is an injective local diffeomorphism; hence
(d) 19 is a diffeomorphism from n onto an open neighborhood U of M.
Now, of the three conditions on p. 71, (1) and (2) are immediate by
construction; hence it only remains to show that, perhaps after shrinking
U, condition (3) also holds true. But this is a local question: it suffices to
prove that
(e) every point x E M has a neighborhood U X C U on which (3) holds.
Indeed, then one replaces U by Ux U X• Thus, fix x E M to prove (e).
Choose a compact neighborhood L of x in M. Let W be the interior of
Lin M, and set e = ~ dist(x, M \ W). Now, let V = W n {liz - xII < e},
and choose 8 < ~e such that L x {liull :s; 8} c n. We claim that UX =
19(V x {liull < 8}) does the job.
Indeed, pick y E U X , that is, p(y) E V and Ily - p(y)11 < 8. Let us
estimate the distance dist(y, M). For z E M \ L, the various choices above
give the following sequence of bounds:
liz - yll 2: liz - xii - Ilx - yll ~ 2e - (11x - p(y)11 + IIp(y) - yll)
> 2e - (E + 8) = e - 8> 8 > Ily - p(y)lI,
74 II. Manifolds

which show that


IIY-ZII > Ily-p(y)11 for Z E M\L and
{
8> Ily - p(y)11 ~ dist(y, M) = dist(y, L).
But L is compact; hence dist(y,L) = Ily - zil for some z E L. This means
that the function h : t H Ily - tl1 2 has a global minimum on M at z, and
consequently, its derivative vanishes on TzM; equivalently, the gradient

gradAh) = 2(y - z)
is perpendicular to TzM, and v = y - z E TzMl.. We have Ilvll = Ily - zll =
dist(y, L) < 8, so that (z, v) E L x {liull :=s: 8} c n. Since {} is injective on
nand z + v = p(y) + (y - p(y)), we conclude that z = p(y) and

dist(y,M) = dist(y,L) = Ily - zll = Ily - p(y)ll·


All of this also shows that p(y) is the unique point z E M such that lIy- z II =
lIy - p(y)lI; hence (3) characterizes p. 0

If one copies the above proof for differentiable manifolds of finite class
C the set liM is a Cr - 1 manifold (the gradients of the local Cr equations
r,

are of class Cr - 1 ), and we end up with a Cr - 1 retraction. But notice


the way in which we apply the Inverse Function Theorem on liM, which
requires r - 1 ~ 1, that is, r ~ 2. These restrictions were avoided in 11.4.1,
p.67, by purely topological means, but there no information concerning
perpendicularity and distances was obtained. That information will be
essential for the Poincare-Hopf Index Theorem.

Exercises and problems


Number 1. Let M C lR m + 1 be a differentiable hypersurface and consider a point c off
M. Prove that a E M is a critical point ofthe Euclidean distance M -+ lR : x H dist(x, c)
if and only if a - c.l TxM.

Number 2. Let 1 : lRP -+ lRn be a smooth map such that the origin is a regular value of
I;hence M = 1-1(0) is a boundaryless smooth manifold of dimension m = p - n. Prove
that there is a diffeomorphism V' : Al x lR n -+ vM such that {) 0 V'(x, u) = x.

Number 3. Let M C lRP be a boundaryless smooth manifold of dimension m. Show


that
N = {(x,u,v) E lRP x lR P x lR P : x E M,u.l TxM, v E TxM}
is a boundary less smooth manifold diffeomorphic to M x lRP .

Number 4. Let M C lRP be a compact boundaryless smooth manifold of dimension m.


For c > 0 we denote
veM = {(x,u) E vM: lIuli < c}.
5. Approximation and homotopy 75

Show that:
(1) For c small enough, iJ restricts to a diffeomorphism from veAf onto an open
neighborhood of A! in lRP .
(2) There is a diffeomorphism h from vA! onto an open neighborhood of A! in lRP
such that h(x,O) = x for every x E A!.
Number 5. Let A! C lR P be a boundaryless smooth manifold of dimension m. For every
continuous mapping c : A! -+ lR everywhere > 0 we denote
veAl = {(x, u) E vA!: Ilull < c(x)}.
Show that:
(1) There is an c such that iJ restricts to a diffeomorphism from veAl onto an open
neighborhood of A! in lR P •
(2) There is a diffeomorphism h from vA! onto an open neighborhood of A! in lRP
such that h(x,O) = x for every x E A!.

5. Approximation and homotopy


In this section we show that homotopies can always be made differentiable,
which is essential for studying homotopies of continuous mappings using
differentiable methods.
As is well known, a continuous homotopy is a continuous mapping
H : [0,1] x M --+ N (of course, one can use any closed interval), and
as customary, we usually denote Ht(x) instead of H(t,x). Thus a homo-
topy is a continuous uniparametric family of mappings H t : M --+ N. Two
mappings f, g : M --+ N are homotopic if there is a continuous homotopy
Ht with Ho = f and HI = g. As is well known, this is an equivalence rela-
tion, whose set of classes is denoted [M, N]. Here there is other standard
notation: the k-th homotopy group is 7rk(N) = [§k, N] for N connected,
and the k-th cohomotopy group is 7rk(M) = [M, §k] (this is indeed a group,
because M is a manifold). This notion is enough for studying continuous
mappings on compact manifolds, but here we will transfer the compactness
assumption from manifolds to maps, which requires some additional care.

(5.1) Proper mappings. A continuous mapping f : M --+ N is called


proper when it is closed and the inverse image of every point of Y is com-
pact. In the context of manifolds, these two conditions are equivalent to
the single condition that the inverse image of every compact set is compact.
In our setting, we can also formulate properness in terms of convergency:
f is proper if and only if every sequence Xk whose image f(Xk) converges
has some convergent subsequence.
Note also that if M is compact, then all continuous mappings are proper.
76 II. Manifolds

As mentioned before, for proper mappings plain homotopy is not conve-


nient. Instead, we consider the notion of a proper homotopy, which simply
is a homotopy H : [0, 1] x M ---t N that is a proper mapping. Then, two
proper mappings f, 9 : M ---t N are properly homotopic if there is a proper
homotopy H t with Ho = f and HI = g.

The preceding definitions can be used in the differentiable setting, just


by requiring all mappings involved to be differentiable. But then, we need
two facts: (i) that every proper mapping is homotopic to a proper differen-
tiable one and (ii) that two differentiable mappings that are homotopic are
homotopic by a differentiable homotopy. To prove this, we must use ap-
proximation techniques and these techniques require the use of retractions
(described in the preceding section).
Theorem 5.2. Let X c R.P be a locally closed set, N c R.q a boundaryless
differentiable CT manifold, and f : X ---t N a continuous (proper) mapping.
For every positive continuous function c : X ---t R. there is a (proper) CT
mapping 9 : X ---t N such that
Ilf(x) - g(x)11 < c(x) for every x E X.

Proof. Since X is closed in some open neighborhood W C R.P, f extends to


a continuous mapping, which we still denote f : W ---t R.q, but the target
need not be N (Tietze Extension Theorem). In the same way, c extends to
a continuous function c : W ---t R., and substituting {c > O} for W, we can
suppose c is positive on W. We proceed now in several steps.
Step I: Approximation of f : W ---t R.q. Since f is continuous, every point
x E W has a neighborhood W X C W such that Ilf(z) - f(x)11 < c(z) for
all z E W X. Let {Ox} be a smooth partition of unity for the covering {WX}
of W, and define

x
Clearly, h is smooth, and for every z E W we have

Ilh(z) - f(z)11 = IIL:f(x)Ox(z) - (L:Ox(z))f(z)11


x x
~ L: Ox(z)llf(x) - f(z)1I < c(z)
x

(notice that if Ilf(x) - f(z)11 2 c(z), then z ¢ W X , and so Ox(z) = 0).


Step II: Retraction of the approximation to N. Let p : U ---t N be a CT
retraction onto N from an open set U C R.q (H.4.l, p. 67). The function
5. Approximation and homotopy 77

c'(x) = dist (f (x) , Rq \ U) is continuous and positive on X. Thus, we can


replace W by {c' > O}, or merely assume f(W) c U. Using min{c,c'}
instead of c in Step I, we get h(W) cU.
For each x E X, let Kx be a compact neighborhood of x in Wand let
Cx > 0 be the minimum of con Kx. The set
Vx = {y E U: IIp(y) - f(x)11 < ~cx}

is an open neighborhood of f(x) = p(f(x)) in Rq and contains a ball cen-


tered at f(x) with radius c5x < CX. Let Bx be the ball centered at f(x)
with radius ~c5x. Now, by continuity, in an open neighborhood AX c
f-I(Bx) n Kx of x in W it follows that Ilf - f(x)1I < ~c5x. By replac-
ing W with the union of the sets A x, we can assume they cover W, and
using a smooth partition of unity {(x}x for the cover {AX}x, we define

c" = L ~c5x(x : W --+ R,


x

which is positive and continuous. Next, we substitute min{c,c',c"} for c in


the preceding step and claim that the composite mapping 9 = po h, which
is well defined and Cr because h(W) c U, is the approximation we sought.
Indeed, let us estimate IIg - fll on X. Pick z E X, and we will have
(x(z) =I 0 exactly for finitely many points x = Xl, ... , Xs E X. Then
s
c"(z) = L ~c5Xi(Xi(Z) ::; ~c5Xll
i=l

c5XI being, to simplify notation, the biggest c5Xi . As z E AX I , f(z) E B XI ;


hence

Ilh(z) - f(xI)1I ::; Ilh(z) - f(z) II + IIf(z) - f(xI)1I < c"(z) + ~c5XI ::; c5Xll

so that h(z) E VXI and IIp(h(z)) - f(XI)11 < ~CXI. On the other hand,
z E AXI also implies Ilf(z) - f(XI)11 < ~CXI' so that

IIg(z) - f(z)11 ::; IIp(h(z)) - f(XI)11 + Ilf(z) - f(XI)11 < cXI·


To conclude, notice that cXI ::; c(z), because z E AXI C K XI .
Step III: Properness. Suppose f proper. The concern then is to choose c
small enough. Pick two locally finite open coverings {Vi} and {Ui} of N such
that Vi C Ui and Vi is compact. Then the compact sets Ki = f-I(Vi) form
78 II. Manifolds

a locally finite covering of X. Hence, each point x E X has a neighborhood


Wx that meets finitely many Ki's, and we denote

Ex = m~ndist(Vi' N \
t
Ud
for the indices i with Wx n Ki =I 0. There is a partition of unity {T/x} for
the covering {Wx }, and define the positive continuous function

E*(Z) = ! L T/x(Z)Ex.
x

We claim that if the approximation 9 is constructed using E ::; E*, then 9 is


proper.
The key fact is that for such a 9 we have g(Ki) C Ui for all i's. Indeed,
suppose Z E Ki = f- 1 (V i ). Since T/x(z) =I 0 implies Z E W x , we see that i
is one of the indices in the definition of Ex, so that Ex ::; dist(V i ,N \ Ui ).
It follows that

Ilf(z) - g(z)11 < E*(Z) ::; ! (L T/x(z)) dist(V i ,N \ Ui) = ! dist(Vi ,N \ Ui).
x

As f(z) E Vi, necessarily g(z) E Ui , as desired.


Finally we deduce from this that 9 is proper. Let LeN be compact.
As the family {Ud is locally finite, L meets finitely many Ui's. But then,
by the property just proved, C = g-1(L) meets only finitely many Ki'S.
As the Ki'S cover X, the closed set C is contained in the union of finitely
many of them. Such a union is compact; hence C is compact too. D

Approximation settles homotopy matters, because:


Proposition 5.3. Let X C RP be a locally closed set, N C Rq a bound-
aryless differentiable CT manifold, and f : X ~ N a continuous (proper)
mapping. There is a positive continuous function E : X ~ R such that
every continuous mapping g : X ~ N with IIf(x) - g(x)1I < E(X) for all
x E X is (properly) homotopic to f.

Proof. To start with, note that any continuous mapping 9 : X ~ N C Rq


is homotopic to f in Rq, by Ht = (1 - t)f + tg. But this homotopy can
map points off N. Then consider a tubular cr
retraction p : U ~ N, and
set c(x) = dist(f(x), Rq \ U) > o. If IIf(x) - g(x) II < E(X) for all x E X, we
deduce
IIf(x) - Ht(x) I = Ilf(x) - (tf(x) + (1- t)g(x)) II
= 11 - tlllf(x) - g(x) I < E(X) = dist(f(x), R n \ U),
5. Approximation and homotopy 79

so that Ht(x) E U. Consequently, we can define a homotopy in N by


po H t . That po H t is proper when f is follows by applying the method of
Step III in the proof of the preceding result 11.5.2 to the proper mapping
[0,1] x X ---+ N: (t,x) t-+ f(x). 0

In the last proof, if f and g are Cr , so is the homotopy obtained. But in


fact, we have the following more general result, essential in the differentiable
setting:

Proposition 5.4. Let X c jRP be a locally closed set and let N c jRq
be a boundaryless differentiable Cr manifold. If two Cr mappings f, g :
X ---+ N are (properly) homotopic, then they are homotopic by a (proper)
Cr homotopy.

Proof. Let H : [0,1] x X ---+ N be a (proper) homotopy with Ho = f,


HI = g. By approximation (11.5.2, p. 76), there is a (proper) Cr mapping
H' : [0,1] x X ---+ N with IIH - H'li arbitrarily small. We thus have a
(proper) Cr homotopy H:,
but we only know that Ilf - Hbll and Ilg - H~II
are small. Anyway, 11.5.3, p. 78, provides (proper) Cr homotopies F t from
f to Hb and G t from H~ to g. Clearly, we can paste Ft to H: and then
with G t to get another homotopy from f to g. The only difficulty is for
the pasted homotopy to be Cr at the junctions. But this can be arranged
easily: replace Ft , H:' and G t by F8(t) , H~(t), G 8(t), for any bump function
o: [0, 1] ---+ [0, 1] with

{ O(t) = ° for t ::; i and


O(t) = 1 for t ~ ~.

These modifications flatten the homotopies near t


the pasting cr.
°
= and t = 1 and make
0

Exercises and problems

Number 1. Let M and N be two differentiable Cr manifolds. Two Cr mappings f, 9 :


M -+ N are cr homotopic if there is a Cr homotopy H : [0,1] x M -+ N such that Ho = f
and Hl = g. Show that this is an equivalence relation for cr mappings. State and prove
the analogous result for proper Cr mappings.

Number 2. Let f, 9 : M -+ lR be two continuous functions such that f (x) < g( x) for
every x E M. Prove that there is a cr function h : M -+ lR such that f(x) < h(x) < g(x)
for every x E AI.
80 II. Manifolds

Number 3. Let M C RP be a differentiable cr manifold and C c M a closed subset. Let


f : M -+ R be a continuous function whose restriction fie : C -+ R is cr differentiable.
Show that for every real number c > 0, there is a CT· mapping 9 : !vI -+ R such that
glc = fie and IIg(x) - f(x) II < c.
Use this result to prove that if !vI is connected, each two points in M can be connected
by a differentiable cr curve.
Number 4. Let H : [a,lJ x M -+ R be a Cr homotopy, where M is a compact differen-
tiable Cr manifold. Show that if a E R is a regular value of H Q , then it is a regular value
of H t for every t > a small enough.

Number 5. Let M be a differentiable cr manifold and let h : M -+ R be a Cr function,


whose zero set we denote by X. Let f : M -+ R be a continuous function that vanishes
on X, and let c : M -+ R be a continuous strictly positive function. Show that there is
a cr function 9 : M -+ R such that IIf(x) - g(x)h(x)1I < c(x) for every x E M.

6. Diffeotopies
Here we discuss homogeneity, which is the property that points in manifolds
can be moved around on demand. To start with, we introduce the key
definition:

Definition 6.1. Let M be a differentiable Cr manifold. A Cr diffeotopyof


M is a differentiable Cr homotopy F : [0, 1] x M -+ M such that Fo = IdJl,[
°
and all mappings Ft , ~ t ~ 1, are Cr diffeomorphisms of M. We say that
F joins each two points x E M and Fl (x) .
We say that F is the identity off a set A C M if Ft(x) = x for x tJ. A
°
and ~ t ~ 1.

Remarks 6.2. (1) Every diffeotopy is a proper homotopy.


Indeed, the mapping P : (t, x) I--t (t, Ft(x)) is a diffeomorphism of [0,1] x
M (use the Inverse Mapping Theorem to check that p- 1 is differentiable) ,'\
hence a proper mapping. Now the projection 7r : [0,1] x M -+ M is proper,)
because [0,1] is compact, and we conclude that F = 7r 0 F is proper too. /
(2) Since every point x E M is connected to Ft(x) by the path t I--t
Ft(x), they lie in the same connected component. Hence, the Ft's preserve
connected components. This allows many reductions to the case where M
is connected.
(3) As for homotopies, the unit interval can be replaced by any other
closed interval. 0
6. Diffeotopies 81

Example 6.3. The mapping (t, x) M tx gives a diffeotopy of JRm on any


closed interval [1, AJ, and it can be transported to the sphere §m c JRm +1
via the stereographic projection from the north pole to obtain a diffeotopy
with equations:

It is clear that Ft moves every parallel Xm+1 = h > -1 to another Xm+1 =


ht, with limt-too h t ---+ 1. Thus we can move the set Xm+1 2:: h into
Xm+1 2:: h' for any given heights h, h'. We say that Ft collapses the sphere
towards the north pole. 0

The main goal of this section is to move points around using diffeotopies.
We start in Euclidean spaces.

°
Proposition 6.4. There is a positive radius E > such that in JRm the
origin can be joined to every point in the open ball Ilxll < E by a smooth
diffeotopy that is the identity off Ilxll 2:: 2.

Proof. Let {} : JR ---+ [0,1] be a bump function such that

{}(o) =1 and
{
(}(t) ==° on It I 2:: 1.

Let a 2:: 1 be such that I{}I(t)1 ~ a for all t E JR, and take E = 1/a i; J.
Let c be a point with a = Ilcll < E. After a rotation we can assume
c = (a,O, ... ,O) E JR x JRm-l, and we consider coordinates x = (y,z) in
JR x JRm-l. We will produce a diffeotopy F joining the origin to c.
Let T : JRm- I ---+ [0,1] be another bump function which is == 1 on Ilzll ~ 1
°
and == on Ilzll 2:: v3. Then define

Ft(x) = Ft(y, z) = (y + t{}(y)T(z)a, z).


It is clear that Fo is the identity and FI (0) = c. Furthermore, if Ilxll =
Vy2 + IIzll2 2:: 2, it must be either y2 2:: 1 or IIzl12 2:: 3; hence either
°
(}(y) = or T(Z) = 0, and Ft(x) = x. Now, we are to prove that each Ft is
a diffeomorphism. To that end, we claim:
For every fixed t, z, the function Ft,z : JR ---+ JR : y M Y + t{}(Y)T{z)a is
bijective.
82 II. Manifolds

Indeed, the function Ft,z is strictly increasing, because its derivative is

1 + to'(y)r(z)a 2: 1 - aa > 0,

i and F t ,z is not bounded, because

1
IFt,z(Y) I 2: Iyl- ItO(y)r(z)al 2: Iyl - -a ~ 00

when Iyl ~ 00, This proves the claim.


It immediately follows from the claim that Ft is bijective; hence to
prove it is a diffeomorphism, we can apply the Inverse Mapping Theorem
and check that the derivative dx (Ft) : jRm ~ jRm is a linear isomorphism,
which can be readily seen by computing the Jacobian determinant:

1 + to'(y)r(z)a *
1 o
= 1 + to'(y)r(z)a > O.
o
o 1

The proof is finished. o


Using this Euclidean construction, the result follows easily for mani-
folds:

Theorem 6.5. Let M be a boundaryless differentiable cr manifold. Let


a, b be two points of M and let A be a connected open set containing both.
Then there is a cr diffeotopy joining a and b, which is the identity off a
compact set K c A, which is a neighborhood of a and b.

Proof. Choose for each point x E A a coordinate domain Ux c A, dif-


feomorphic to jRm. By II.6.4, p. 81, we find two open relatively compact
neighborhoods Wx C Vx of x such that Kx = V x C Ux and each point of
Wx can be joined to x by a diffeotopy of Ux which is the identity off Kx;
notice that this latter condition means that the diffeotopy can be extended
to the whole manifold M by the identity. Now, since A is connected and
covered by the Wx's, we find a finite chain W X1 , ... , W Xp with a E W X1
and b E W Xr ' Then pick points Yk E W Xk n WXk+l' k = 1, ... ,p - 1, and
set Yo = a, YP = b. By construction, for each k = 0, ... ,p - 1, we can
join Yk and Yk+l by a diffeotopy F(k) of M which is the identity off the
compact set KXk+l cA. Clearly, if we paste those diffeotopies, we get one
which joins a = Yo and b = yp and which is the identity off the compact set
6. Diifeotopies 83

K = U KXk+l C A. The only problem, that the pasting be Cr at the junc-


tions, is arranged by a bump function () as in 11.5.4, p. 79. The composite
G * F of two diffeotopies F and G is

(G F)( ) {F(()(2t), x) for t ~ !,


* t,x = FI(G(()(2t _ 1),x)) for t ~ !. D

Corollary 6.6. Let M be a boundaryless differentiable Cr manifold of di-


mension m ~ 2. Let al, ... ,ap and bI, ... ,bp be two collections of points of
M, and let A be a connected open set containing both. Then there is a Cr
diffeotopy joining ak and bk for all k = 1, ... ,p, which is the identity off a
compact set K c A.

Proof. By induction on p. Assume the result for fewer that p points, and let
N = M \ {a p , bp }. By assumption, there is a diffeotopy of N joining ak and
bk for k = 1, ... ,p-l, which is the identity off a compact set LeA \ { ap , bp }
(which is connected because dim (A) ~ 2); such a diffeotopy extends to M.
Then, by 11.6.5 above, there is a diffeotopy of M joining a p and bp that is
the identity off a compact set K contained in the connected set A \ {ak' bk :
1 ~ k < p}. One concludes by pasting these two diffeotopies. D

Remarks 6.7. (1) On M = li every diffeomorphism is either decreasing or


increasing, which immediately gives a restriction to joining series of points:
they must be numbered in exactly the same (or the reverse) order. Then it
is easy to produce by hand the diffeotopy that connects the two series (or
to do the pasting as in the proof above). This is the essential restriction
for manifolds of dimension 1.
(2) An easy consequence of the preceding result is that connected man-
ifolds are n-homogeneous (in particular homogeneous): any two collections
of n points can be transformed to each other by a diffeomorphism (in di-
mension 1 with the restriction explained in the above remark).
(3) Another interesting consequence that often makes life easier is that
any finite collection of points of a connected manifold has an open neigh-
borhood diffeomorphic to lim.
Indeed, given Xl, ... , X p , pick an open set U diffeomorphic to lim, and
choose any YI, ... , Yp E U. By remark (2), there is a diffeomorphism f,
mapping Xk t--t Yk; hence f-I(U) is the open neighborhood we sought. D
84 II. Manifolds

All these results obviously work for bordered manifolds as long as the
points involved are not in the boundary. We will not need an elaborate
discussion.

Exercises and problems


Number 1. Let f : IRm -+ IRm be a linear isomorphism with positive determinant. Show
that there is a smooth diffeotopy F : [0,1] x IRm -+ IRm such that Ft = f.
Number 2. Let h : IR -+ [0,1] be a smooth function with h(t) = 1 for It I ~ 1 and
h(t) = 0 for It I ~ 4. Prove that there exists a positive real number E > 0 such that for
every a E IRm with Iiall < E and every linear mapping u : IRm -+ IRm with II IdRm -ull < E,
the map
Ft(x) = x + th(lIxIl2)(U(X) - X + a)
is a smooth diffeotopy of IRm verifying (i) Ft(x) = x for 0 ~ t ~ 1 and Ilxll ~ 2 and
(ii) Fl(X) = u(x) + a for Ilxll ~ 1. (Check that each Ft is a proper local diffeomorphism
whose fibers all have the same number of points.)
Number 3. Let f : IRm -+ IR m be a smooth diffeomorphism with f(O) = 0 and dof =
IdRm. Show that there is a smooth diffeotopy F: [0,1] x IRm -+ IRm such that Ft = f.
Number 4. Let U be an open neighborhood of the origin in IRm, and let f : U -+ IRm
be a differentiable Cr mapping with f(O) = 0, dof = IdRm. Consider the function h in
Problem Number 2, and prove that there is a positive real number E > 0 such that

Ft(x) = x + th( lI~f )(f(x) - x)


is a well-defined Cr diffeotopy of IR m verifying (i) Ft(x) = x for 0 ~ t ~ 1 and Ilxll ~ 2E
and (ii) Fl(X) = f(x) for IIxll ~ E.
Number 5. Let sm C IRm+1 be a sphere of odd dimension m. Construct a diffeotopy
F ofSm such that F1 (x) = -x for all x E sm.
Number 6. Let F: [0,1] x sm -+ sm be a diffeotopy.
(1) Show that

tP: [0,1] x U -+ U = IRm +1 \ {O} : (t, x) I-t IlxllFt (II~II)


is a diffeotopy.
(2) Pick any point a E sm. Show that the Jacobian determinant at a of tPt is positive.
Use this to prove that if m is even, there is no isotopy of sm such that Fl(X) = -x
for all x E sm.

7. Orientation
In this final section of the chapter we quickly survey the notion of orienta-
tion and some related constructions.
7. Orientation 85

(7.1) Orientation of differentiable manifolds. An orientation ( on


a linear space E of finite dimension m 2: 1 is a choice of a basis B =
{UI, ... , urn}, two choices being equivalent if the determinant of the base
change matrix is positive; we denote ( = [UI, ... , urn] and say that B is
a positive basis of (. Clearly, there are only two orientations, which are
denoted by ( and -( and are called opposite. In E = jRm there is a
canonical orientation (rn, corresponding to the canonical basis {el' ... ,ern}.
We know the two first examples of this very well: in jR2, the counterclockwise
orientation; in jR3, the screwdriver orientation.
Now, let M be a differentiable manifold of dimension m. An orientation
on M is a family (M = {(x : x E M} of orientations (x in each tangent
space TxM, such that each point a E M has a coordinate system x on
whose domain U

[a~llx'···' a:rn Ix] = (x for every x E U.

When this condition holds, the coordinate system (and the corresponding
parametrization) is said to be compatible with the orientation.
Notice here that the base change matrix for two bases of partial deriva-
tives is the Jacobian matrix of the corresponding change of coordinates;
hence the Jacobian determinant of the change of coordinates of a pair of
compatible parametrizations is positive (we just say that the change is posi-
tive). By definition, if M has an orientation, it has a covering by coordinate
domains with positive changes, which is called a positive atlas. Conversely,
if there is a positive atlas, we can use (*) above to define an orientation (
onM.
Suppose M is connected and oriented by (. Pick a coordinate domain
U, with coordinate system x. To check condition (*) for x, we must use
a coordinate system compatible with ( and check whether the change is
positive. By continuity, this holds (or not) locally; hence, (*) holds (or not)
locally. If the coordinate domain is connected, (*) holds (or not) on the
whole domain. Combining this with the fact that any two points of Mean
be joined by a chain of connected coordinate systems, we conclude that once
we know ( at a point, then we know it everywhere. In other words, M has
exactly two different orientations, ( and its opposite: -( = {-(x : x E M}.
Finally, let us mention here that the product M x N of two oriented
manifolds is immediately oriented at (x, y) E M x N by
86 II. Manifolds

given that
(x = [UI,"" urn] and (y = [VI, ... , vn].
In short, we write (MxN = ((M, (N).

(7.2) Orientation and differentiable mappings. Consider a differen-


tiable mapping of oriented manifolds of the same dimension, f : M -+ N.
Let x E M be a regular point of f, so that dxf : TxM -+ Tf(x)N is a linear
isomorphism. Then we say that f preserves (resp., reverses) orientation at
x E M if dxf maps a positive basis of (M,x onto a positive basis of (N,/(x)
(resp., -(N,/(x»' We denote

signxU) = +1 (resp., -1)


and call this the sign of f at x.

The sign can be easily computed by localization through coordinates


compatible with both orientations (M and (N: as the Jacobian matrix
of such a localization is the matrix of dxf with respect to the bases of
partial derivatives, one only has to check whether that Jacobian has positive
determinant. In particular, we see that if f : M -+ N preserves (resp.,
reverses) orientation at x, it does the same on a whole neighborhood of x.
In particular, if a diffeomorphism of connected manifolds preserves (resp.,
reverses) orientation at some point, it does so at all of them. The following
is an application of this:

Proposition 7.3. Let Ft be a diffeotopy of an oriented manifold M. Then


all Ft's preserve orientation.

Proof. Since diffeotopies preserve connected components, we can suppose


!vI connected, and then it is enough to check that the Ft's preserve the
orientation at a given point a E M. To see that, let t E [0,1] and choose
a local parametrization 'I/J : V -+ M compatible with the orientation of
M, such that 'I/J(O) = Ft(a). Then F-I('I/J(V» is a neighborhood of (t, a),
and there are (i) a local parametrization rp : U -+ M compatible with the
orientation of M, with a = rp(O), and (ii) a neighborhood J c I of t, such
that F(J x rp(U» c 'I/J(V). In this situation, the mapping

(J' : J -+ lR : s f--t det do ('I/J -1 0 Fs 0 rp)

is well defined and continuous. Since 'I/J, rp, and the Fs's are diffeomor-
phisms, (J' never vanishes and has constant sign on a neighborhood of t. In
conclusion, for s close enough to t, Fs preserves or reverses orientation as Ft
7. Orientation 87

does. It follows that all Ft preserve or reverse orientation simultaneously.


But Fa = IdM preserves; hence all Ft preserve also. 0

(7.4) Orientation of hypersurfaces. Let M c ]R.m+1 be a differentiable


hypersurface, that is, dim(M) = m. We look for a normal vector field
v : M -+ ]R.m+1, which is a mapping such that 0 =lv(x) 1.. TxM for every
xEM.
To start with, fix a coordinate domain U corresponding to a parametriza-
tion cp. Then for every x E U, v(x) must be perpendicular to the partial
derivatives

a
~I = ~(cp -1 (x)), ... , acpm+1
(aCP1 a (cp -1)
(x)) ,1 ~ z.~ m.
UXi x UXi Xi

This perpendicular vector is easily obtained as a vector product. Take new


variables u = (U1,"" Um +1) and let

A(u) =

Then, let the i-th component of v(x) be the signed adjoint to Ui in A(u).
Computing detA(u) for u = a'!i Ix
(1 ~ i ~ m) through its first column, it
follows immediately that this v(x) is a normal vector field on U, which can
be assumed unitary after dividing by its norm.
Now, notice that at each point x E M there are only two unitary vectors
perpendicular to TxM. This makes it reasonable to paste together the v's
obtained from different coordinate systems. For this, we must understand
when v(x) = v'(x) for two different parametrizations. But by construction

(
m+ 1 _
-
[a a ] _ [, a a]
v, aX1 ' ... 'aXm - v, ax~ , ... , ax'm

is the canonical orientation in ]R.m+1, and it follows readily that v(x) = v'(x)
if and only if the change of coordinates is positive. From this we see that M
has a global normal vector field if and only if M is orientable. Furthermore,
in that case, there are exactly two different unitary normal vector fields,
opposite each other. We can distinguish them by chosing an orientation (:
v is compatible with ( if the coordinate systems in (*) above are compatible
with (.
88 II. Manifolds

Examples 7.5. (1) Let Me jRm+l be a closed differentiable hypersurface,


which has a global equation M : f = 0, that is, a differentiable function
°
f : jRm+l ~ jR such that is a regular value of f and M = f- 1 (0). Then,
M disconnects jRm+l: otherwise, f would have constant sign; hence every
point x E M would be extremal, hence a critical point of f!
Once f is given, we know (11.2.8, p. 57) that each tangent space TxM
is perpendicular to the gradient gradx(f) = (/t(x), ... , axa.!+1 (x)), which
is always 1= 0, and consequently is a global normal vector field (and 11 =
grad(f)/II grad (f) II is unitary). Thus, Mis orientable.
In fact, this is another formulation of the Jordan Separation Theorem
for differentiable hypersurfaces: every closed differentiable hypersurface of
a Euclidean space disconnects it, has a global equation, and is orientable.
We will prove this using degree theory (111.6.2, p. 124, 111.6.4, p. 129)
(2) The first case to which we apply the above construction is the unit
sphere §m C jRm+l, whose equation is I:i x~ = 1. Then the gradient
1I(x) = x is already unitary and gives the canonical orientation ( of §m.
We always equip §m with this orientation; 11 is called the outward normal
vector field.
Let us now look at two antipodal points a and -a in §m. The two
tangent spaces Ta§m and T _a§m coincide: both are perpendicular to the
line through a and -a. Thus we can compare the orientations (a and (-a:

and
(-a = [VI, ... , vmJ if (m+l = [-a, V1,···, vmJ.
Thus we can take V1 = -UI, V2 = U2, .•. , Vm = Um and conclude that
(-a = -(a.
We have just seen that the sphere has opposite orientations at antipodal
points.
(3) Now suppose a = (0, ... ,0,1) and -a = (0, ... ,0, -1). Let 7fN be
the stereographic projection from a and let 'Irs be that from -a (11.2.4, p.
55). We have

(a = [e1, ... , emJ if and only if m is even.

After this remark, from 11.2.9, p. 58, we see that 'Irs preserves orientation if
and only if m is even. Since (-a = -(a, we similarly see that 'lrN preserves
orientation if and only if m is odd. Hence, 'Irs and 'lrN do not form a
7. Orientation 89

positive atlas. To get one, change the sign of one coordinate in one of the
projections. 0

(7.6) Orientation of inverse images. Let f : M -+ N be a differentiable


cr mapping of oriented manifolds, and let a E N \ aN be a regular value of
both f and flaM. As we know, f-l(a) is a manifold (II.3.2, p. 62), and we
can use the orientations of M and N to define an orientation on f- l (a).
We do it as follows.
Let x E f-l(a). By II.3.2, p.62, L = Txf-l(a) is the kernel of the
derivative dxf; hence this restricts to a linear isomorphism E -+ TaN for
any linear supplement E C TxM of L. Choose any such E with a basis
VI, ... , Vn that dxf maps onto a positive basis WI, ... , Wn of TaN. Then we
define an orientation ~x on L by declaring positive any basis UI, ... , U m - n
of L such that VI, ... ,Vn , UI, ... ,Um - n is a positive basis of TxM.
We must see that this does not depend on the choice of data, so assume
other v~, uj are given. Consider the matrix A = ( ~l 112 )
of the base
change from {vLuj} to {Vi,Uj} and the matrix C = (CliO) of dxf with
respect to the bases {Vi, Uj} and {Wi}. In view of the way orientations
determine the choices, we have the following:

(i) The matrix C 1 is the base change from {d xf (Vi)} to {wd, so 0 <
det(CI ).

(ii) Since CA = (CIAIIO), the matrix CIA I is the base change from
{dxf(v~)} to {Wi}; hence 0 < det(CI ) det(A I ), and by (i), 0 < det(Ad.

(iii) The matrix A is the base change from {v~, uj} to {Vi, Uj }; hence 0 <
det(A) = det(AI) det(A 2 ), and by (ii), 0 < det(A 2 ).

But A2 is the base change from {uj} to {Uj}, and consequently {uj} defines
the same orientation as {Uj}.
Since we know that the definition does not depend on choices, let us
see why ~ = {~x : x E f-l(a)} is indeed a well-defined orientation on
our inverse image. To do this, pick coordinates (XI, ... , xm) of M and
(YI, ... ,Yn) of N as in the proof of II.3.2, p. 63, so that (Xl' ... ' X m- n ) are
coordinates of f- l (a) and
90 II. Manifolds

Thus we can use the vectors Vi = ox m-n+. 0 .1, x Wi = ayOI ~.I x


• a ' and Uj = VA] .
It is quite clear that depending on how x, y respect orientations, either
(Xl, ... ,Xm - n ) or (Xl, ... , -x m - n ) defines ~ consistently at all points of
f-l(a) close to x. We are done.

(7.7) Orientation of boundaries. The definition of orientation applies


of course to manifolds with boundary, as no boundaryless assumption has
been made as far. If M has boundary 8M, all tangent spaces TxM carry
an orientation, despite whether or not x belongs to 8M, In fact, it is easy
to see that M is orientable if and only if the boundaryless manifold M \ 8M
is; in other words, the orientation of M at boundary points is determined
by the nearby interior points. A more interesting matter is that when M
is oriented, its boundary can be oriented as a manifold itself. We describe
this here.
Let M be a manifold with boundary 8M, and fix an orientation (on M.
Choose any point a E 8M. The tangent space Ta (8M) is a hyperplane in
TaM, and we pick a coordinate system at a, x : U -+ lliIm with Xl 2 O. Then
the linear isomorphism dax : U f--t t = (tl' t2,'" ,tm ) maps Ta (8M) c TaM
onto {O} x IR m- 1 c IRm. Thus we find two types of vectors that are tangent
to M but not to 8M: the inward vectors, with tl > 0, and the outward
vectors, with tl < O. As usual, via changes of coordinates, one sees that
these conditions do not depend on the choice of the coordinates, and we
have a consistent notion of outward tangency.
Curve germs provide a more geometric way to determine inward and
outward vectors:

(1) If we have "( : [0, c) -+ M with "((0) = a E 8M and "('(0) = U E


TaM \ Ta(8M), then U is inward.

(2) If we have "( : (1 - c, 1] -+ M with "((1) = a E 8M and "('(I) = U E


TaM \ Ta(8M), then U is outward.

- X
7. Orientation 91

In (1), Xl 0, : [0, c) -t [0,00) must be increasing, and tl = (Xl 0,)'(0)


is positive; hence U is inward. In (2) we get an outward vector, because
Xl 0 , : (1 - E, 1] -t [0,00) is decreasing.

Finally, we define an orientation on Tx(8M) by

8(= [u]
(= [v,u]

Again, some straightforward computations confirm that this indeed defines


an orientation 8( on 8M.

Examples 7.S. (1) The orientation of ~m = ~m X {O} as the boundary


of lHIm+l : Xl 2: 0 is the "wrong" one! Indeed, lHIm+1 carries the canonical
orientation (m+l, but since el is inward tangent, we get 8(m+l = _(m.

-el = outward

(2) Let us look at the unit sphere §m as the boundary of the ball L:i x~ ~
1. Then the vector v(x) = X is outward tangent to the ball, and we see
that the orientation of the sphere as boundary of the ball is the canonical
orientation defined in 11.7.5(2), p. 88.
(3) A case of relevance for later discussion is that of a cylinder. Let M be
an oriented boundaryless manifold and consider the cylinder M' = [0,1] xM
oriented as described in 11.7.1, p. 85: if UI, ••. , U m give the orientation (x
at X EM, at (t, x) E M' take the orientation

~(t,x) = [(1,0), (0, UI), .•• , (0, um)].


92 II. Manifolds

This ~ induces an orientation 8~ on 8M', which is a disjoint union of two


copies of M:
8M'=MoUM1 ,
'Po :Mo = {O} xM == M: (O,x) == X,
'PI :Ml = {l}xM == M: (l,x) == x.
Then 'PI preserves orientations and 'Po reverses them. In other words, the
orientation 8~ induced on 8M' is the right one on Ml and the wrong one
on Mo. The picture below describes the situation.

• •
t
~ • [0,1]

Indeed, it is enough to remark that (1,0) is inward at (0, x) and outward


at (1, x). Moreover, this is immediate, looking at the curve germs

{
1'0 ~ [0,1) -+ M:~ t r-t (t, x), 1'0(0) = (0, x), 1'~(0) = (1,0),
1'1 . (0,1] -+ M . t r-t (t, x), 1'1 (1) = (1, x), 1'1 (1) = (1,0). 0

Exercises and problems

Number 1. Let M and N be diferentiable Cr manifolds of dimensions m and n, respec-


tively, M boundaryless. Let (M and (N be orientations on M and N, respectively.
(1) Equip M x Nand N x M with the orientations (MXN and (NXM, and determine
when the canonical diffeomorphism T : M x N -7 N x M : (x, y) >--+ (y, x) preserves
orientation.
(2) Note that M x aN = o(M x N) and compare the orientations (Mx8N and O(MXN.
Number 2. Let f : jRm -7 jRm be a diffeomorphism. Show that there is a diffeotopy
F: [0,1] x ]Rm -7 ]Rm with F[ = f if and only if f preserves orientation.

Number 3. Show the following orient ability criterion. A manifold M is orientable if


and only iffor any orientations (u, (von two connected open sets U, V C M and for any
two points x, y E Un V, (u,x = (v,x if and only if (u,y = (v,y.
7. Orientation 93

Number 4. Use the criterion of the preceding problem to show that:


(1) The Mobius band M C R3 parametrized by

x = cosO + t cos 0 cos !, y = sinO + tsinOcos!, z = tsin!


is not orientable.
(2) The real projective plane is not orientable.
Number 5. Show that real projective spaces RJlDTn are orientable exactly for odd m.
What about complex projective spaces?
Number 6. Let M C RP be a boundaryless smooth manifold of codimension n with a
smooth equation 1 : RP -t Rn (that is, M = 1-1(0) and 0 is a regular value of f). Prove
that M is orientable. Find another sufficient condition for orientability involving 11M.
Number 7. Find the orientation of the cilinder M C R3 : x 2 + y2 = 1, 0 ::; z::; 1 such
that the local diffeomorphism R2 -t M : (t, x) t-+ (cos t, sin t, z) preserves orientation.
Then determine the orientation induced on each connected component Co, C 1 of the
boundary aM:

and study whether the mapping Co -t C 1 : (x, y, 0) t-+ (x, y, 1) preserves orientation or
not.
Number 8. Let M be a boundaryless differentiable Cr manifold of dimension m. Prove
the following:
(1) Let a, bE M be two points in a connected open set A c M and let.>. : TaM -t
nM be a linear isomorphism, preserving orientation il M is oriented. Then there is a
C diffeotopy joining a and b, which is the identity off a compact neighborhood K C A
of both a and b and such that da F1 = .>..
(2) If M is oriented and connected and 'P, t/J : B -t Ml are two parametrizations
compatible with the given orientation, both defined on the unit open ball BeRm, then
there is a Cr diffeotopy F : [0,1] x M -t M such that F1 o'P = t/J on some smaller open
ball B' c B.
Chapter III

The Brouwer-Kronecker degree

Here we present the beautiful Brouwer-Kronecker degree theory for arbitrary proper
mappings from the differentiable viewpoint. To start with, §§1 and 3 contain the
construction of the Brouwer-Kronecker degree. Also, we use it to prove the Funda-
mental Theorem of Algebra. In §2 we take a small diversion to describe the degree
through integration by means of the de Rham cohomology. Next, we discuss in §4
the extension of the theory to arbitrary differentiable manifolds, which is based on
the computation of the degree of a C1 mapping. In §5 we define the so-called Hopf
invariant, for mappings between spheres of unequal dimension, which is based
upon the notion of link coefficient. We devote §6 to a beautiful application of the
theory: the Jordan Separation Theorem. We show that every closed differentiable
hypersurface disconnects the Euclidean space, and from this we deduce that such a
hypersurface has a global equation and is orientable. We conclude the chapter by
proving in §7 the famous Brouwer Fixed Point Theorem, jointly with some inter-
esting consequences, and the equally important theorem by Brouwer concerning
the existence on spheres of vector fields without zeros.

1. The degree of a smooth mapping


Here we present the notion of degree for smooth data. The first step towards
the definition of the Brouwer-Kronecker degree is as follows:
Proposition 1.1. Let f : M --+ N be a proper smooth mapping of two ori-
ented, boundaryless, smooth manifolds of dimension m. Then each regular
value a E RJ C N of f has an open neighborhood V C RJ such that for
every b E V, the inverse image f-l(b) is a finite set and the integer

d(f, b) = L signx(f)
xEJ-1(b)

only depends on a.

Proof. First note that for any regular value b the compact set f-l(b) is
discrete. This follows because f is a local diffeomorphism at each x E f-l(b)
95 -
96 III. The Brouwer-Kronecker degree

by the Inverse Mapping Theorem. We must now show that b t---+ d(f, b) is
locally constant in Rf (which is open because f is proper). To that end,
we need an open neighborhood V of a such that f-l(V) is a disjoint union
of finitely many connected open sets Uk such that each restriction fJUk is
a diffeomorphism onto V.
To obtain V, let f- l (a) = {Xl, ... , x r }. By the Inverse Mapping The-
orem, there are disjoint open neighborhoods Uk of Xk, such that fJu~ is
a diffeomorphism onto some open neighborhood V' of a, which we can
take common for all the k's. Then since f is closed, the set f(M \ Uk Uk)
is closed. As that set does not contain a, we can find a connected open
neighborhood V C V' of a such that

V n f(M \ UUk) = 0.
k

Hence, f-l(V) C Uk Uk' and we take Uk = f-l(V) n Uk' so that f-l(V) =


Uk Uk·
Note also that, Uk being connected, signx(f) is constant on Uk, say
== (Tk. Thus, for every bE V we have

d(f, a) = L signxk (f) = L (Tk = L signYk (f) = d(f, b).


k k k D

Proposition 1.2 (Boundary Theorem). Let X be an oriented smooth man-


ifold of dimension m + 1 with boundary aX = Y and let N be an oriented,
boundaryless, smooth manifold of dimension m. Let H : X -t N be a
proper smooth mapping and let a be a regular value of Hand HJy. Then

d(HJy,a) = o.
Proof. Denote f = HJy, which is a proper smooth mapping, because Y is
closed in X. By II.3.2, p. 62, the inverse image C = H-l(a) is a compact
smooth curve with boundary

aC = C n Y = f-l(a).
1. The degree of a smooth mapping 97

By the classification theorem for compact curves (II.3.3, p. 63), the finitely
many connected components rk of e are either circles (without bound-
ary points) or arcs with two boundary points Pk, qk E 1-1 (a). Thus, the
boundary points of Be come in pairs, and we have

d(f,a) = Lsignpk(f) + Lsignqk(f).


k k

We claim that signpk (f) = - signqk (f), which concludes the argument.
Henceforth we omit the indices k in all notation. The situation is depicted
in the figure below.

')" (1)

-')" (0)

We consider in the component under consideration, r, the orienta-


tion it has as an inverse image (II.7.6, p.89). Let t t--+ ')'(t) E r be a
parametrization such that ')'(0) = p, ')'(1) = q. After renaming p,q as q,p
and reparametrizing by t t--+ ')'(1- t), we can suppose that the inverse image
orientation in T'Y(t)r is in fact given by the tangent vector ')"(t). We next
discuss orientations at the end points of r.
(i) Orientations at p. Let Ul, ... , Urn be a positive basis of TpY. As Y is
oriented as the boundary of X, v, UI, . .. , Urn is a positive basis of TpX for
any outward tangent vector v. But ')"(0) is inward (II.7.7, p. 90); hence we
can take v = -1"(0). Thus

-,),'(0), UI, ... , Urn

is a positive basis ofTpX. On the other hand, we know that 1"(0) gives the
orientation of r as an inverse image; hence
98 III. The Brouwer-Kronecker degree

is a positive basis of TpX if and only if dpf maps Ul, ... , U m onto a positive
basis of TaN, that is, if and only if signp(f) = +1. But the determinant of
the change from the first base to the second is (-1) m+l, and so
signp(f) = +1 if and only if m is odd.
(ii) Orientations at q. The argument is the same, but at q the tangent
vector 'Y'(I) is outward; hence we do not need -'Y'(I), and the sign of the
base change is (_1)m. Thus:
signq (f) = +1 if and only if m is even.
From these two equivalences, we see that the signs signp(f) and signq(f)
are opposite, as desired.
This concludes the proof of the proposition. D

Next we consider smooth homotopies:


Proposition 1.3 (Homotopy invariance). Let H : [0,1] x M -+ N be a
proper smooth homotopy of oriented, boundaryless, smooth manifolds, and
let a be a regular value of both HfJ and HI. Then
d(Ho, a) = d(Hl' a).

Proof. By the Sard-Brown Theorem, II.3.4, p. 63, we can choose regular


values b of H arbitrarily close to a, and then

d(Ho, a) = d(Ho, b), d(Hl' a) = d(Hl' b)


by IILl.I, p. 95. We apply the preceding proposition with X = [0,1] x M,
so that Y = {O} x M U {I} x M, and get d(Hly, b) = O. But, since Y is a
union of two disjoint components,

d(Hly, b) = d(HI{o}XM, b)+d(HI{l}xM, b).


We can identify
{O} x M == M and HI{o}XM == Ho,
{I} x M == M and HI{l}xM == HI
and then see what the degrees are. But we must care for the orientations
induced on Y as a boundary. Indeed, as we know (II.7.8, p. 91), in a cylinder
like X the orientations induced in {O} x M and {I} x M are reversed;
hence if, say, the equivalence {O} x M == M preserves orientations, then
{I} x M == M does not, and we have
d(HI{o}xM, b) = d(Ho, b) and d(HI{I}XM, b) = -d(HI' b).
1. The degree of a smooth mapping 99

Substituting in (*), we see that d(Ho, b) = d(Hl' b). Clearly, if {O} xM == M


reverses orientations, the conclusion is the same. 0

Proposition and Definition 1.4. Let f : M -t N be a proper smooth


mapping of two oriented, boundaryless, smooth manifolds of dimension m;
furthermore, N is connected. Then the integer d(f, a) does not depend on
the choice of the regular value a E RI' It is called the degree of f and is
denoted by deg(f).

Proof. Let b be another regular value of f. Since N is connected, by 11.6.5,


p. 82, there is a smooth diffeotopy Ft of N with Fl(b) = a. Then F t 0 f is
a proper smooth homotopy of f and Fl 0 f. Since a is a regular value of
these two mappings, from III.1.3, p. 98, we get

d(f, a) = d(Fl 0 f, a).

But Fl is a diffeomorphism that preserves orientation everywhere (II.7.3,


p. 86); hence by the chain rule:

d(Flof,a) =
o

Remarks 1.5. (1) It is clear that the above definition depends on the
chosen orientations. The essential thing here is that if the orientation of M
is reversed, then the degree changes sign.
(2) It is also clear from the definition that every non-surjective smooth
mapping has degree zero. 0

Examples 1.6. Let §m C ~m+l denote the standard unit sphere.


(1) Of course the identity has degree 1. Now let us look at the antipodal
map f : x t--+ -x. It is a diffeomorphism. Hence its degree is ±1 according
to the behavior of orientations, which can be checked at any point. Pick the
north pole a = (0, ... ,0,1). Then f(a) = -a = (0, ... ,0, -1) is the south
pole, and the two tangent hyperplanes Ta§m and T _a§m coincide as linear
spaces, but they carry opposite orientations. Now, daf is the symmetry:
u t--+ -u, which maps any basis {Ul, ... ,um } to {-Ul, ... ,-um }. The
determinant of the latter with respect to the former is (-l)m, that is:

{ +1, and so the bases define equal orientations, if m is even,


-1, and so the bases define opposite orientations, if m is odd.
100 III. The Brouwer-Kronecker degree

Hence, a positive basis of Ta§m is mapped to

a negative basis of T _a§m if m is even,


{
a positive basis of T_a§m if m is odd,

and we conclude that daf

reverses orientation if m is even,


{
preserves orientation if m is odd.

All in all, deg(f) = -1 if m is even, and deg(f) = +1 if m is odd.


(2) To produce a mapping with degree -1 in any dimension, we turn to
symmetries, like h : x t---+ (Xl, ... , X m , -X m +1)' This h maps again the north
to the south pole, but here dah is the identity, because the tangent vectors
have their last component zero. Consequently, a discussion as above shows
that dah : Ta§m --t T_a§m reverses orientations, and deg(h) = -1.
(3) We can identify the unit circle §l C R2 with the group of complex
numbers of module 1. Then the mapping §l --t §l : z --t zd (complex
multiplication) has degree d.
If d = 0, this mapping is constant and its degree is indeed O. Hence, we
assume d t= O. In polar coordinates, we can represent the mapping by

z = cosO + isinO t---+ zd = cos(d·O) + isin(d·O).


Clearly, there are no critical values, and each value has Idl preimages. To
check orientations, just look at the representation

§l --t §l: (cosO,sinO) t---+ (cos(d.O),sin(d.O)),

which preserves orientation for d > 0 and reverses it for d < O. In fact, we
have just formalized the idea of winding §l around itself. We will generalize
this later (V.1.1, p. 183). This example shows quite clearly why degree is
called degree! 0

The degree of a composite mapping is easy to compute:

Proposition 1. 7. Let f : M --t Nand g : N --t P be two proper smooth


mappings of oriented, boundaryless, smooth manifolds, with Nand P con-
nected. Then
deg(g 0 f) = deg(g) . deg(f).
1. The degree of a smooth mapping 101

Proof. Let a E P be a regular value of 9 0 f. Then every y E g-l(a) is a


regular value of f. Indeed, let x E f-l(y). As a is a regular value of 9 0 f,
the derivative dx(g 0 f) is an isomorphism. But by the Chain Rule,

and we conclude that dxf is injective, hence an isomorphism, because all


spaces have dimension m. Consequently, we have

deg(f) = L signx(f) for every y E g-l(a).


XE/- 1(y)

Thus

deg(g 0 f) = L signx(g f) = L L 0 signy (g) signx(f)


xE(gof)-l(a) yEy-l(a) xE/-1(y)

= L signy(g) L signx(f) = L signy(g) deg(f)


yEg-l(a) XE/-1(y) YEy-l(a)

= deg(g) . deg(f).
o
Now we prove two fundamental properties of deg: .

Proposition 1.8 (Boundary Theorem). Let X be an oriented smooth man-


ifold of dimension m+ 1 with boundary ax = Y, and let N be a connected,
oriented, boundaryless, smooth manifold of dimension m. Let H : X -+ N
be a proper smooth mapping. Then

deg(Hly) = o.

Proof. The smooth mapping Hly is proper, because Y is closed in X. Then,


by the Sard-Brown Theorem, we can pick a point a E N that is a regular
value of both Hand HIY. Hence

deg(Hly) = d(Hly,a),
and the result follows readily from 111.1.2, p. 96. o
Proposition 1.9. Let H : [0,1] x M -+ N be a proper smooth homotopy
of oriented, boundaryless, smooth manifolds, N connected. Then

deg(Ho) = deg(Hl)'
102 III. The Brouwer-Kronecker degree

Proof. By the Sard-Brown Theorem, the two proper smooth mappings Ho


and HI have some common regular value a E N, and then

deg(Ho) = d(Ho, a), deg(HI) = d(HI' a).


From this and III.1.3, p. 98, we get deg(Ho) = deg(HI)' D

We end this section with a proof of the Fundamental Theorem of Alge-


bra, which in the first chapter was presented as the origin of degree theory.
The proof that follows brings in the preceding notions and makes clear
the interest of developing the theory for proper mappings. Later we will
give a second proof of this result, closer to Gauss's ideas, appealing to the
Euclidean degree (IV.2.7, p. 150)
Proposition 1.10 (Fundamental Theorem of Algebra). Every algebraic
equation zP + CIZp-1 + ... + Cp = 0 with complex coefficients has some
complex solution.

Proof. The mapping

p:]R2 == C --+ C: (x, y) == x + iy = z H zP + qzp-I + ... + cp


is smooth and proper (because limlzl-too IP(z)1 = 00). By III.1.5(2), p. 99,
it is enough to show that deg(P) i= O. To do that, consider the homotopy

Pt(z) = zP + tCIZp-1 + ... + tcp , 0:::; t :::; 1,


which is proper and smooth. Hence deg(P) = deg(Po). But the degree
of the mapping Po : z H zP is exactly p. Indeed, one checks immediately
that (i) any a i= 0 is a regular value of Po with p different roots c in C and
(ii) at each of them Po preserves orientation: dcPo is multiplication by the
complex number pcp-I. D

Exercises and problems


Number 1. For each i = 1, ... , r, consider a proper smooth mapping Ii : Ali ~ Ni with
dim(Mi) = dim(Ni ), Ni connected. Prove that deg(/1 x ... x Ir) = deg(fI)··· deg(fr).
Deduce that 011 the torus T m = §1 X ... X §1 there are smooth mappings T m ~ T m of
arbitrary degree.

Number 2. Let p, q, ml, ... ,mr be even positive integers, and let n = p + q, m =
L:i mi .
(1) Show that the following smooth mapping has degree 2:
2. The de Rham definition 103

(2) Define by induction a mapping fr : §ml X ... X §mr ---+ §m of degree 2r-l, hence
not null-homotopic.

Number 3. Check that the following quadratic mapping is well defined, and compute
its degree:

Number 4. Let m = 2n + 1 be an odd positive integer.


(1) Compute the degree of the canonical surjection
11": §m ---+ JR]pm : (xo, ... ,Xm) 1-+ (xo : ... : Xm).
(2) Show that the mapping h of the preceding problem factorizes through 11", and
obtain a smooth mapping f : JR]pm ---+ §m with deg(f) = l.
(3) Take a different view of f via stereographic projections.
Number 5. Fix an orientation on C m == R2m. Let U C C m be an open set, and let
f : U ---+ C m be a proper holomorphic function. Show that deg(f) is the cardinal of the
generic fiber of f: the inverse image f-l(a) of every regular value a of f has exactly
deg(f) points. In particular, deg(f) ~ O.
Number 6. It can be shown that the degree of a smooth mapping CIP'2 ---+ CIP'2 must be
a perfect square. Construct all of them.

Number 7. Fix an orientation on the complex projective space ClP'm to compute the
degree of the mapping
ClP'm ---+ ClP'm : z = (zo : ... : Zm) 1-+ Z = (zo : ... : Zm),
where Zi stands for the complex conjugate of Zi E C.

Number 8. Given a proper smooth mapping f : M ---+ N of smooth manifolds with


boundary, N connected, such that f(aM) c aN, define

deg(f) = L signx(f),
xE!-l (a)

where a is a regular value of f off aN. Develop the corresponding theory, using homo-
topies H t such that Ht(aM) c aN. Prove also that deg(f) = deg(flp), where P C aM
is the inverse image f-l(Q) n aM of any chosen connected component Q of aN.

2. The de Rham definition


In this section we describe the Brouwer-Kronecker degree of a smooth map-
ping in terms of differential forms and integration. We will not fully depict
all details behind the constructions, only those directly linked to the notion
of degree.

(2.1) Forms and de Rham cohomology. Let M be a connected, ori-


ented, boundaryless, smooth manifold of dimension m. We denote by
104 III. The Brouwer-Kronecker degree

r~(M) the linear space of all smooth differential forms of degree k, k-


forms for short, with compact support. For k = 0, r~(M) consists of all
smooth functions f : M -+ JR. with compact support. As is well known, the
derivatives of such a function f give its total differential df : x t-+ dxf, and
this induces a linear operator d : r~(M) -+ r~(M). Then, this operator
extends to d : r~(M) -+ r~+l(M) for all k, so that dod = 0 and

d( a 1\ ,8) = da 1\ ,8 + (-1 r a 1\ d,8


for every r-form a and s-form ,8; d is called exterior differentiation.
A k-form w with compact support is called closed when its exterior
differential is zero, Le., dw = 0, and exact when w = da for some (k - 1)-
form with compact support a. The closed k-forms with compact support
are a linear subspace of r~(M) and the exact k-forms with compact support
another linear subspace. As dod = 0, exact implies closed, and we can
consider the quotient linear space

Hk(M JR.) = {closed k-forms with compact support}


c' { exact k-forms with compact support} .

This is the so-called k-th de Rham cohomology group of M with compact


support. It is a deep theorem that this cohomology coincides with the
singular cohomology H~(M, Z) after extension of coefficients (H~(M, JR.) =
JR. 0z H~(M,Z)), but we will not discuss this matter here. Of course, if
M is compact, everything has compact support and we get the ordinary
cohomology groups, but in the non-compact case, it is important to con-
sider compact supports. For instance, H:;n(JR.m, JR.) = JR., but disregarding
compact supports in all definitions above, we get Hm(JR.m, JR.) = O.
Here we will concentrate on the maximum degree m and look solely
at the m-th cohomology group H:;n(M, JR.). This group can be beautifully
described by integration. Indeed, since M is boundaryless, Stokes' Theorem
says simply that
1M da=O,
which means that the linear mapping

is well defined. But in fact, it can be shown that it is an isomorphism.


By some standard reduction using partitions of unity and local coordinate
systems, this amounts to proving the following assertion:
2. The de RhaIn definition 105

Let h : IRm -? IR be a smooth function with compact support, such


that IIRn h = O. Then

for some smooth functions hI' ... ' h m with compact support.
In other words, h is the divergence of some vector field with compact
support. This is proved in parametric form, by induction on m, using
elementary integration.

(2.2) Homomorphisms on cohomology. Let M and N be two con-


nected, oriented, boundaryless, smooth manifolds of dimension m. Every
proper smooth mapping f : M -? N induces linear mappings

j* : r~(N) -? r~(M) : w t---+ j*w.

These pull-back mappings are compatible with all operations with forms,
including exterior differentiation, so that they induce homomorphisms on
the cohomology groups, which we still denote f*. Again, we concentrate
on the maximum degree group. Then, we have a commutative diagram of
linear mappings
H:-(N,IR) ~
r H:-(M,IR)

L 1~ ~ lL
~ IR

Thus we are ready to prove the main result concerning degree and co-
homology.
Theorem 2.3. Let f : M -? N be a proper smooth mapping of connected,
oriented, boundaryless, smooth manifolds of dimension m. Then there is
an integer d such that
fMj*W=d·Lw
for every m-form w with compact support on N. Moreover, for any regular
value a E N of f we have

d = L signx (f) (0 if f-I(a) = 0).


xEf-l(a)

Proof. The mapping oX in the preceding commutative diagram is a linear


mapping IR -? 1R; hence it must be multiplication by some real number 6,
106 III. The Brouwer-Kronecker degree

and commutativity reads:

for every m-form w with compact support on N. Then, by II.3.4, p. 63, we


can pick a regular value a of f, and the theorem follows at once if we show
that
6= 2: signxU)
xEf-l(a)

(note that the right-hand side is indeed an integer). In other words, it is


enough to see that

1J*w = 2:
M xEf-l(a)
signA!) 1w
N

for a suitable w that we construct now.


First of all, since a is a regular value, its inverse image f- l (a) is finite,
suppose non-empty, say f- l (a) = {Xl, ... , x r }. Actually, as in the proof of
IILl.1, p. 95, we find an open neigborhood V of a in Nand r disjoint open
neighborhoods Ul , •.. , Ur of Xl, ... , xr in M, such that the restrictions
fk = fluk : Uk --+ V are diffeomorphisms and f-l(V) = U1 U ... U Ur .
Finally, we shrink V to a domain of coordinates diffeomorphic to jRm.
Next, using any diffeomorphism V --+ jRm, we pull back to V the m-form
<p( X )dXl /\ ... /\ dX m ,

where <p : jRm --+ [0, 1] is a smooth bump function such that

<p(x) = {1 for II xII ~ 1,


o for Ilxll 2: 2.
This construction guarantees that the pull back can be extended by zero
to the whole manifold N and that it has compact support. We denote that
extension by w, and we will check the equality (*) for it.
On the one hand, since the support of w is contained in V, we have

r w = 1vr w = ± 1~m
1N
r <p(X)dXl ... dX m =f= 0

(change of variables for the difeomorphism V --+ jRm). On the other hand,
the support of the form J*w is contained in f-l(V) = U Uk, and since the
Uk'S are disjoint, we have

r J*w = 2:1
1M k Uk
J*w.
2. The de Rham definition 107

Here we can compute each summand using the corresponding diffeomor-


phism fluk : Uk ---t V, and again by the change of variables for integrals,
we get

1Uk
f*w = Uk f
Jv
W,

where Uk = ±1 according to whether or not fluk preserves or reverses ori-


entation (which holds on the whole Uk because this open set is connected).
In other words, since Xk E Uk,

and we conclude that

as desired. Note that this determines the integer L:k sign xk (f), because
fN w 1= O.
The reader will easily supply the simplified argument that settles the
case f-l(a) = 0. 0

Thus, the de Rham approach gives the following most general form of
the change of variables for integrals:

Corollary 2.4. Let f : M ---t N be a proper smooth mapping of connected,


oriented, boundaryless, smooth manifolds of dimension m. Then

1M f*w = deg (f)·1 w

for every m-form w on N with compact support. In particular, the integral


is zero if f is not surjective

Of course, this statement contains the classical change of variables of


calculus, when f is a diffeomorphism.
On the other hand, the invariance of homotopy follows quite easily in
this context:

Corollary 2.5. Let M, N be two connected, oriented, boundaryless, smooth


manifolds of dimension m, and let H : [0,1] x M ---t N be a proper smooth
homotopy. Then
deg(Ho) = deg(Hl)'
108 III. The Brouwer-Kronecker degree

Proof. Let w be an m-form with compact support on N, and consider H*w.


By the properties of exterior diferentiation we have dH*w = H*dw. But
dw = 0, because m = dim(N), so that

o= f H* dw = f dH* w.
J[O,l]XM J[O,l]XM

Now we compute the latter integral using Stokes' Theorem. Since the
boundary of [0,1] x M is a disjoint union of two copies Mo and Ml of M,
with wrong and right orientations, respectively (11.7.8(3), p. 91), we have

o= f dH*w = f H;w - f How.


J[O,l]XM JM JM

Thus,
deg(Ho) L w = 1M How = 1M H;w = deg(Hl) L W;

hence deg(Ho) = deg(Hl). D

Example 2.6. One of the deepest theorems in mathematics, the Gauss-


Bonnet Theorem, can be formulated as the computation of a degree. Let
us describe this.
Fix a connected, closed, boundaryless smooth hypersurface M C Rm+1
of dimension m. Then M is orientable, disconnects R m +1, and has a global
equation f = 0 (all of this is the Jordan Separation Theorem; see 111.6.4,
p. 129). What is relevant here is that

11 = grad(f)/II grad (f) II : M ~ Sm C R m +1

is a unitary global normal field (11.7.5(1), p.88). Replacing f by - f if


needed, 11 is compatible with the orientation chosen on M (11.7.4, p. 87),
and it describes the volume element of M: OM = det(lI, .). This 11 is the
so-called Gauss mapping. Then, at each point x E M we have the derivative

dx ll : TxM ~ Tv(x)sm = hyperplane perpendicular to II(X) = TxM.


This is the Weingarten endomorphism, and it is self-adjoint. Indeed, a
straightforward computation shows that for every u, v E R m +1 perpendic-
ular to II(X), that is, for every u, v E TxM,
1
(v, dx ll(u)) = II grad(f)II (v, dx { grad(f)) (u))
2. The de Rham definition 109

and dx ( grad (f) ) is a self-adjoint endomorphism of Rm+1, its symmetric


matrix being (8!2Jx (x) ) .
j

One can accept that this endomorphism dxv of the tangent space TxM
measures the way M c R m twists at x, and its determinant K{x) =
det{dxv) is the Gauss curvature of M at x. As dxv is self-adjoint, its
eigenvalues are all real and the endomorphism diagonalizes. Those real
eigenvalues are the principal curvatures of M at x, and their product is
precisely K (x).
Next, suppose M is compact. Then the number

/'i,= fMKnM
is well defined, and quite naturally, it is called the integral curvature of M.
Thus we find the integral considered by Dyck in 1888 (I.3, p. 31).
To explore this integral curvature further, consider the volume element
nsm of the sphere. A straightforward computation shows that KnM
v* nsm, and we have

Now the Gauss-Bonnet Theorem says that if m is even,

where X{M) is the Euler characteristic of M. We conclude that the Gauss-


Bonnet Theorem just says:
The degree of the Gauss map of an even-dimensional compact hy-
persurface is one half of its Euler characteristic.
This statement can be proved by means of the Poincare-Hopf Index
Theorem. We will discuss this in V.7. 0

Exercises and problems


Number 1. Compact supports are essential. Show that any smooth form Q = fdxl 1\
... 1\ dXm in IR m has some primitive, but this may well not have compact support.

Number 2. Let 'P : IR -t §l be the local diffeomorphism t t-t (cost,sint). Let Q be a


I-form on §l. Show that:
110 III. The Brouwer-Kronecker degree

(1) There is a periodic smooth function 9 : ~ -+ ~ such that 'P*a gdt, and
I(t) I;
= 9 is a smooth function such that I(t + 271") = I(t) + a. Ist
(2) If Ist a
= 0, I induces a smooth function h: §I -+ ~ with = dh. a
Deduce from the above that the linear map Ist : HI (§I ,~) -+ ~ is indeed injective.
Number 3. Consider the torus T m = §I x .. · X §I. Prove that Hk(Tm,~) =I 0 for
0::; k::; m.
Number 4. Let T = §I X §I.
(1) From each factor in T = §I X §I obtain a linear subspace of HI (T,~) isomorphic
to HI(§I,~).
(2) Use the usual periodic local diffeomorphism 'P : ~2 -+ T to represent the I-forms
a on T by forms Idx + gdy on ~2, with I and 9 periodic, and prove that the linear
mapping

is injective.
Conclude that dimlR(HI(T,~)) = 2, and so HI(§I x §I,~) = ~2.
Number 5. Consider the smooth surface M C ~3 given by x 2 + y2 + Z4 = 1, and
consider the smooth mapping I : M -+ §2 : (x, y, z) H (x, y, z2). Compute the integral
IM /*w for the form w = x 2 dy 1\ dz - y 2 dx 1\ dz + z 2 dx 1\ dy on §2. Is the computation
really necessary?
Number 6. Consider in ~3 the two diffeomorphic surfaces x 2 +y2 = z and X4 +y4 = z.
Check that their Gauss curvatures are, respectively,

K= 4
(4X2 + 4y2 + 1)2
Study the extrema of both, as well as their lines of constant curvature.
Number 7. Let M C ~3 be the torus parametrized by
x = (R + r cos u) cos v, y = (R + r cos u) sin v, z = r sin u.

Then:
(1) Show that the Gauss mapping of M is
vex, y, z) = (- cosucosv, - cosusin v, - sin u).
(2) Obtain the matrix of the Weingarten endomorphism with respect to the basis
a/au, a/av of T(x,y,z)M.
(3) Deduce that K(x, y, z) = :\
r
(1 - v';t
.,2+ y 2).
(4) Compute the integral curvature of M.

3. The degree of a continuous mapping


So far, we have defined the Brouwer-Kronecker degree for proper smooth
mappings only. To have the notion for arbitrary proper mappings, it is
3. Tbe degree of a continuous mapping 111

enough to recall that every proper mapping is homotopic to a proper smooth


mapping by a proper homotopy:
Proposition and Definition 3.1. Let f : M -+ N be a proper mapping of
two oriented, boundaryless, smooth manifolds of dimension m; furthermore,
N is connected. Then all proper smooth mappings 9 : M -+ N properly
homotopic to f have the same degree, and we define the degree of f by
deg(f) = deg(g).

Proof. If g, g' : M -+ N are properly homotopic to f, then they are prop-


erly homotopic themselves, and by II.5.4, p. 79, there is a proper smooth
homotopy H t such that Ho = 9 and HI = g'. But then, deg(g) = deg(g')
by III.1.9, p. 101. That indeed there are smooth mappings 9 homotopic to
f follows from 11.5.2, p. 76, and II.5.3, p. 78. 0

As for smooth mappings, we have the product formula for composite


continuous mappings:
Proposition 3.2. Let f : M -+ Nand 9 : N -+ P be two proper map-
pings of oriented, boundaryless, smooth manifolds, with Nand P connected.
Then
deg(g 0 f) = deg(g) . deg(f).

Proof. Let f' and g' be proper smooth mappings properly homotopic to
f and g, respectively. Then g' 0 f' is a proper smooth mapping properly
homotopic to 9 0 f. Thus:

deg(g 0 f) = deg(g' 0 f') = deg(g') . deg(f') = deg(g) . deg(f). D

We also have the following:


Proposition 3.3 (Boundary Theorem). Let X be an oriented, smooth
manifold of dimension m + 1 with boundary ax = Y, and let N be a
connected, oriented, boundaryless, smooth manifold of dimension m. Let
H : X -+ N be a proper mapping. Then

deg(Hly) = o.

Proof. Let H' be a proper smooth mapping properly homotopic to H. Then


H'ly is a proper smooth mapping properly homotopic to Hly, so that
deg(Hly) = deg(H'ly), and the latter degree is zero by III.1.8, p. 101. 0

Similarly, we deduce invariance by homotopy:


112 III. The Brouwer-Kronecker degree

Proposition 3.4. Let H : [0, 1] x M -+ N be a proper homotopy of oriented,


boundaryless, smooth manifolds, N connected. Then

deg(Ho) = deg(Hd.

Proof. Let H' : [0, 1] x M -+ N be a proper smooth mapping properly


homotopic to H, so that Hb and H~ are proper smooth mappings properly
homotopic to Ho and HI, respectively. We have

deg(Ho) = deg(Hb) and deg(HI) = deg(HD,


and by 111.1.9, p. 101, deg(Hb) = deg(Hi). D

Thus we have an invariant attached to every proper mapping between


two manifolds M and N, the Brouwer-Kronecker degree, which only de-
pends on the homotopy class of the mapping.

Example 3.5. Consider the degree -1 mapping f : §I -+ §I : z 1---7 liz


(this is 111.1.6(3), p. 100, for d = -1). Then define on the torus T = §I x§I
the mapping F = f x f : T -+ T. This map has degree 1, but it is not
homotopic to the identity.
Indeed, to compute deg(F), look at the derivative of F at the fixed
point a = (1,1):

daF(u, v) = (dIJ(u), dIJ(v)) = (-u, -v).

This preserves orientation; hence deg(F) = 1. Next, suppose there is a


homotopy Ht with Ho = F, HI = IdT. Denote by j : §I -+ T the injection
z 1---7 (z, 1) and by p: T -+ §I the projection (z, z') 1---7 z. Then ht = poHtoj
is a homotopy with ho = f and hI = Id~p. But such a homotopy cannot
exist, because deg(f) = -1 and deg(Id§l) = 1.
Thus we see that, in general, degree does not fully classify homotopy
classes. We will see that for spheres it does classify: this is the content of
the famous Hopf Theorems (V.2.1, p. 191, and V.3.1, p. 196). D

Exercises and problems


Number 1. Prove that a proper mapping f : 1R ~ 1R has degree 0, +1, or -1. Show
that the three cases occur. Compute the degree of a real polynomial mapping f(x) =
x d +alxd- 1 + ... + Cd. What does this say about the solutions of the equation f(x) = O?
3. The degree of a continuous mapping 113

Number 2. Construct two proper mappings f,g : IR2 -+ IR2 which are homotopic but
have different degree (of course, they will not be properly homotopic).
Number 3. Prove a product formula for the degree of continuous mappings, like that
of Problem Number 1 of 111.1.
Number 4. Construct a proper mapping f : IR2 -+ IR2 of degree 1 such that f-1({y =
O,x =I O}) is the graph of the topological sinus y = sin(l/x), x =I O. Extend fto §2
via stereographic projection from the north pole, and describe the inverse image by that
extension of the meridians corresponding to the coordinate axes.
Number 5. Fix integers m, n ::::: O. Every point z in §m+n+1 C IR m+ n+2 = IRm+1 x IR n+ 1
can be written uniquely as

z = cos(t)x + sin(t)y, where x E §m, y E §n, 0 ~ t ~ 7r/2,

except that x (resp., y) is not determined for t = 7r/2 (resp., t = 0). Given two continuous
mappings f : §m -+ §m and f : §n -+ §n, their join is defined by

f * 9 : §m+n+1 -+ §ffi+n+1 : z>-+ cos(t)f(x) + sin(t)g(y).


Prove that f * 9 is continuous, and its degree is deg(f) deg(g).
Number 6. Consider the exponential mapping

exp: IR -+ §1 : x >-+ exp(x) = (cos(27rx),sin(27rx)).

Now, let f : §1 -+ §1 be a continuous mapping of degree d, and let h : [0,1] -+ IR be a


continuous lifting of h = f oeXPilO,l] (that is, h = expoh). Show that d = h(l) - h(O).
Number 7. Let f, 9 : §m -+ §m be continuous mappings, m ::::: 2. Check that the
mapping h = - f + 2(f, g}g : §m -+ §ffi is well defined and then compute its degree as
follows:
(1) Reduce to the case that f == (1,0, ... ,0) on the upper hemisphere and 9 ==
(1,0, ... ,0) on the lower one.
(2) Note that h coincides on the upper hemisphere with (2g~ -1, 2g 1 g2, ... , 2g1gm +1)
and compute the degree of the latter in terms of deg(g).
(3) Note that h coincides on the lower hemisphere with (/1, - h, ... , - fm+I) , and
compute the degree of the latter in terms of deg(f).
(4) Conclude that deg(h) = (-l)mdeg(f) + (1- (-l)m)deg(g).
Number 8. Let p,: §ffi X §m -+ §m be a continuous mapping.
(1) Show that for every y E §m all mappings p,(., y) : §m -+ §m have the same degree,
which we denote deg 1(p,).
(2) Define a degree deg 2(p,) similarly.
Now let g, f : §m -+ §m be two continuous mappings, and consider the mapping
p,(f,g) : §m -+ sm. Show that

deg(p,(f,g» = deg(f) deg 1 (p,) + deg(g) deg 2 (p,)·


How does this generalize the previous problem?
114 III. The Brouwer-Kronecker degree

4. The degree of a differentiable mapping


So far, the degree of a mapping f is computed through some smooth ap-
proximation by the formula in IlL 1.4, p. 99. However, it is natural to
expect that in case the mapping f is itself differentiable, the same for-
mula works directly for f, without any interposed smooth approximation.
This is indeed true but, somehow surprisingly, a quite delicate technical
matter. Indeed, a first naive attempt would be to revise all constructions
in I1Ll replacing smooth by differentiable cr mappings. Then, for homo-
topy invariance, we must apply the Sard-Brown Theorem to a Cr mapping
[0, 1] x M -+ N: this requires the difficult version for finite class mappings,
and even more, it needs class r > dim ([0, 1] x M) - dim (N) = 1. Thus
we see the essential difficulty that differentiable Cl mappings are excluded
by this approach. In fact, as we will see, this is the only obstruction to a
Cl degree theory as neat as the smooth one we have previously presented.
This is the part of the theory that relies on some results that we do not
prove, and it is intentionally confined to this section.
After this preamble, we get into the main result:

Theorem 4.1. Let f : M -+ N be a proper Cl mapping of two oriented,


boundaryless, smooth manifolds of dimension m; furthermore, N is con-
nected. Let a E N be a regular value of f. Then f-l(a) is a finite set
and
deg(J) = L signx(J)·
xEf-l(a)

Proof. We will use a close smooth approximation 9 : M -+ N as given by


I1.5.2, p. 76, with some additional conditions. Since a is a regular value and
f is proper, f-l(a) consists of finitely many points, say CI, ... , Cr. Then
we pick r local coordinates that preserve orientations on disjoint domains
Uk == jRm at each point Ck and another coordinate domain V == jRm at a
such that Ck == 0, a == 0, and f(Uk) C V. Denote by fk : jRm -+ jRm the
localizations of f at each Uk. We claim the following

Let ( = fk' and let det(;;: (0)) =I- 0. There is a closed ball B =
{llxll
e:B -+
:S p} such that for
jRm verifies
E > ° small enough, if a smooth mapping

8e · (x) _8(·
_t (x) I <
Ile(x) - ((x) II < E, 1
_t
8xj 8xj
E, for all i,j,

then:
4. The degree of a differentiable mapping 115

(1) 0 = ~(b) for a unique b, and Ilbll < p,


(2) det (;;: (0)) . det (;!: (b)) > 0 (that is, signo(() = signb(~)).
Assume this for the moment. Pick a smooth approximation 9 of f such
that, for every x EM,

IIg(x) - f(x)11 < dist (J(M \ W), a), dist (J(T), N \ V),

where W is the union of the sets Wk == {llxll < p} and T is that of the
sets Tk == {llxll :s: p} (by construction these distances are not zero). These
bounds give:

(i) a tf. g(M \ W). Indeed, if a = g(x) with x tf. W, we have


dist (J(M \ W), a) :s: IIf(x) - g(x)ll,

which is impossible.

(ii) g(Tk) C V. For, if g(x) tf. V with x E Tk, we get

dist (J(T), N \ V) :s: Ilf(x) - g(x)ll,


a contradiction.

Then, the approximation 9 has well-defined localizations gk : B -+ jRm,


and the crucial point here is that 9 can be chosen so that each ~ = gk
verifies the bounds on the derivatives required by the claim. This is the
part we assume without proof.
We also have that g-l(a) C W; hence

g-l(a) == Ugk1(0),
k

and by the conclusions of the claim, there are exactly r points bk E gk 1 (0),
none is critical, and

We conclude that

deg(f) = deg(g) = Lsignbk(gk) = Lsignck(f).


k k

The proof is thus complete, except for the following:


116 III. The Brouwer-Kronecker degree

Proof of the Claim. This is a review of the the proof of the Inverse Mapping
Theorem. Consider the continuous mapping

Since ((0) = a ::f. 0, there is an 'T/ > 0 such that

1 a-det (Z;i. (O)+tij)


J
1 < ~Ial
Then choose p > 0 such that

a(i (Xi) -
1-ax· -a(i (-)1 1
0 <-'T/ (hence I((x) - al < ~Ial),
J ax·J 2

for II xIiI , ... , IIxmll < p.


Next, denote by 8 > 0 the minimum of 1I((x)1I on the sphere IIxll = p
(this minimum is not zero, because by construction 0 is the unique zero of
(). Then, choose E: > 0 smaller than ~1J and ~8. We will check that this E:
does the job.
Define [ mimicking (. Hence

- _ a~i (_ ) a(i (-)


~(x) = det (a(i - )
ax. (0) + tij , where tij = -a Xi - -a 0,
Xj Xj
J

and on IIXili < p we have

so that by (*), la - [(x)1 < ~Ial, which clearly implies a· [(x) > OJ hence
a~· ~
a· det ( axt. (x)) > 0 for IIxll < p.
J

This already guarantees (2) in the statement of the Claim, as soon as b is


given.
To find it, note that the b we sought is a minimum of the function
IIxll = p we
J..L(x) = 1I~(x)1I2 in the compact set B. But on its boundary
have

1I~(x)1I ~ 1I((x)II-II~(x) - ((x) II > 8- E: > E: > II~(O) - ((0)11 = II~(O)II,


4. The degree of a differentiable mapping 117

which means that the minimum does not belong to the boundary. Hence,
b is a critical point of the smooth mapping /-L on the open ball Ilxll < p, so
that its gradient is zero at that point. But a simple computation gives

gradb(/-L) = 2~(b) (:!i. J


(b»),

and since the determinant of the latter matrix is not zero, we conclude
~(b)= 0 as desired.
Finally, we turn to the uniqueness of b. Suppose that ~(x) = 0 with
IIxll < p. We apply the Mean Value Theorem to each component ~i on the
segment [b, xl and find points Xi E [b, x], hence IIxili < p, such that

o = ~i(X) - ~i(b) = L. [)Xj


[)~i (Xi)(Xj - bj ).
J

This is a linear system in the unknowns Zj = Xj - bj, and its determinant is


~(x) -=1= 0 (see above). Hence the only solution is Zj = 0, that is, x = b. 0

(4.2) Degree theory for differentiable manifolds. Once we know


that computations are the same for differentiable as for smooth mappings,
the theory goes through for differentiable manifolds. As explained before
(II.2.7, p.57), the key result needed is smoothness: every differentiable
Cr manifold is Cr diffeomorphic to a smooth manifold. Such a smooth
manifold is constructed by approximation, in a more delicate sense than
the one developed in II.5.2, p. 76. Actually, it is the kind of approximation
also used without proof for III.4.1, p. 114: approximation of derivatives.
As mentioned already, here is where we put the limits on our presentation.
Now, we sketch the construction of the Brouwer-Kronecker degree in
the differentiable setting. Let f : M --+ N be a proper C1 mapping of
oriented, boundaryless, C1 manifolds of dimension m, N connected. Then
there are C1 diffeomorphisms r.p : M --+ M' and 'IjJ : N --+ N' onto smooth
manifolds M' and N'; we choose in these smooth models the orientations
compatible with r.p and 'IjJ and define
deg(J) = deg('IjJ 0 f 0 r.p-l).

Of course, this deg(J) does not depend on the diffeomorphisms r.p and 'IjJ.
To prove it, pick any regular value a E N of f. Then, from the preceding
theorem, III.4.1, p. 114, it easily follows that

deg(J) = L signx(J),
xEf-l(a)
118 III. The Brouwer-Kronecker degree

which clearly does not depend on 'P or 'ljJ.


After this, one extends the theory to C1 manifolds using smooth models:
proper mappings are properly homotopic to proper C1 mappings whose
degree we can compute by the sum of signs formula, the Boundary Theorem
holds, and it follows that properly homotopic mappings have the same
degree, and the degree of a composite mapping is the product of the degrees
of the factors. All arguments follow without surprise.

Exercises and problems

Number 1. Let ( : R m -+ R m be a differentiable mapping with ((0) = 0 such that do(


is a linear isomorphism. Show that for p > 0 small enough

IIdoc 10 (d z ( - do() I <~ for Ilzll < p.


Pick a positive real number 8 < II doC 111- 1. Prove that for TJ > 0 small enough any
smooth mapping ~ : R m -+ R m such that

verifies the two conditions

Use these bounds and the Finite Increment Theorem to deduce that ~ has a unique
zero c with Ilell < p and the signs at c of the Jacobian determinants of ~ and ( coincide.
Number 2. Show that the mapping f : §1 -+ §1 defined by
f(x, y) = (cos(27rN), sin(27rN))
is C1 but not C2 , and compute its degree.

Number 3. Let f : M -+ N be a proper mapping of two oriented, boundaryless, smooth


manifolds of dimension m, N connected. Suppose that there is an a E N such that f is
smooth on some neighborhood U of f-1(a) and a is a regular value of flu. Prove that
there is a close smooth approximation 9 : M -+ N of f that coincides with f on U (up
to shrinking) and such that g-1(a) C U. Deduce that

deg(f) = L signx(f)·
xEf-l(a)

What is the difference between this and III.4.1, p. 1147


Number 4. Compute the degrees of the following two mappings §1 -+ §1:

() (X2_y2,-2XY) forx~O,
1 9 ( x,y ) = {
(X 2 _y2,2xy) for x:S; o.

( 2) h(x )_ {(X' 2-y) for x ~ 0,


,y - ((x _ 3y 2)X, (3x 2 - y2)y) for x :s; O.
5. The HopE invariant 119

Number 5. Complete the details of C1 mapping degree theory as suggested in 111.4.2,


p.117.
Number 6. Let M C 1R2 denote the graph of the function h: IR ~ IR given by

h(x) = {o X 3/ 2
for x ::; 0,
for x ~ 0.

Prove that AI is a C1 curve, but not C2 • Fix an orientation on !vI and find proper
mappings f : IR ~ M of all possible degrees.

5. The Hopf invariant


Although the notion of degree is mainly intended for discussing mappings
of manifolds of the same dimension, it can be profitably used to treat some
mappings between spheres of different dimensions. The pioneering ideas in
this respect are due to Hopf (see 1.3, p. 30).
We fix the following notation. Consider the north and south poles,
p = (0, ... ,0,1) and -p = (0, ... ,0, -1), of the unit sphere §2m-l C 1R2m ,
m 2: 2, and let ¢ : §2m-l \ {-p} --+ Tp§2m-l == 1R2m - 1 be the stereographic
projection from the south pole. Let n c §2m-l X §2m-l be the open set
defined by the inequalities x =f. y, x =f. -p, y =f. -p, and consider the smooth
mapping
2m-2 ( ) ¢(y) - ¢(x)
iP: n -+ § : x, y t-t 1I¢(y) _ ¢(x)11 .

We have the following:


Proposition and Definition 5.1. Let f §2m-l --+ §m be a smooth
mappzng.

(1) If f is not surjective, we let H(f) = 0.


(2) Iff is surjective, consider any two distinct regular values a, b =f. f( -p)
of f. Then f-l(a) x f-l(b) c n is a compact, oriented, boundaryless,
smooth manifold of dimension 2m - 2, and iP restricts to a smooth
mappzng
¢a,b : f-l(a) x f-l(b) -+ §2m-2,
whose degree does not depend on the choice of a and b. We let

H (f) = deg( ¢a,b)'

The integer H (f) is called the Hopf invariant of f.


120 III. The Brouwer-Kronecker degree

Proof. Suppose f surjective. By II.3.2, p. 62, and 11.7.6, p. 89, the inverse
image of every regular value is an oriented manifold of dimension m - 1;
hence the product f-1(a) x f-1(b) is an oriented manifold of dimension
2m - 2. On the other hand, since the compact sets f-1(a) and f-1(b) are
disjoint and do not contain the south pole -p, their product is contained
in D. All in all, the degree deg(4)a,b) is well defined.
Next, let us see why that degree does not depend on a (the proof for
b would be the same). Let a' =1= f( -p) be another regular value of f, and
let Ft be a diffeotopy of §m that fixes band f (-p) and is the identity off a
neighborhood W of a, a' such that F1 (a') = a. We are to apply a boundary
argument twice to conclude that deg(4)a,b) = deg(4)a l ,b).
Suppose first that a' is close to a. Then we take We Rj, which implies
that a is a regular value of the smooth mapping H(t, x) = Ft(f(x)). Indeed,
if Ht(x) = a, we have Ft(f(x)) = a, and f(x) must be in W, so that x is a
regular point of f. Hence dxf is surjective, and since Ft is a diffeomorphism,
the derivative
dxHt = dj(x)Ft 0 dxf
is surjective. With this settled, H-1(a) is a compact smooth manifold with
-p (j. Ht-1(a) and Ht-1(a) n f-1(b) = 0 for all t.

For the first condition, notice that if a = Ht ( -p) = Ft(f( -p)), then
f( -p) E W, a contradiction. For the second, if a = Ht(x) = Ft(f(x)),
then f(x) E Wand b (j. W.
Thus the smooth mapping

Ai X H- 1() f-1(b) ~2m-2 ( ) 4>(y) - 4>(x)


~: = a x --t.:l) : t,x,y I-t 114>(y) _ 4>(x) II
is well defined, and by the Boundary Theorem (III.l.8, p. 101)
0= deg(4)lax), where ax = aH-1(a) x f-1(b).
Note that the boundary of H-1(a) consists of the two pieces {O} x f-1(a)
and {I} X f- 1 (a'), one with the right and the other with the wrong orien-
tation. Thus ax consists of
{O} x f-1(a) x f-1(b) and {I} x f-1(a') x f-1(b),

one with the right and the other with the wrong orientation. Clearly, 4>
restricts to 4>a,b on the first component and to 4>a' ,b on the second, and by
the remark concerning orientations we conclude

0= deg(4)lax) = deg(4)a,b) - deg(4)a l ,b).


5. The HopE invariant 121

Once this case is proved, assume a' is arbitrary. Then choose a dif-
feotopy Ft as above, but now a need not be a regular value of H t = Ft 0 f.
Then, by the Sard-Brown Theorem, there is a regular value al of H close
to a so that Fl1(al) = a~ is close to a'. By the case already proved, we
have
deg(cPa,b) = deg(cPal,b) and deg(cPa',b) = deg(cPa~,b).
On the other hand, the argument above with H-l(aI) instead of H-l(a)
shows that
deg(cPal,b) = deg(cPa~,b)'
and, as desired, we deduce

deg(cPa,b) = deg(cPa',b). o

Remarks 5.2. (1) The degree of the mapping cPa,b is in fact the link coeffi-
cient of the two manifolds f-l(a) and f-l(b) as was introduced by Brouwer
(see 1.2, p. 28). We use the stereographic projection to set the two manifolds
in a Euclidean space. We will abuse notation and write

although this disregards the sign on the left-hand side; for our use here this
is only a notation.
(2) After the above definition of the Hopf invariant, it is only natural to
ask whether for non-surjective mappings we can also compute it through
link coefficients. The answer is that we can. Namely, if f : §2m-l -t §m is
not surjective and a, b =1= f ( -p) are two regular values such that f- 1 (a) =1= 0
and f-l(b) =1= 0, then

To see this, just repeat the proof of II1.5.1, p. 119, by choosing a second
regular value a' ~ f(§2m-l). Then f-l(a') = 0 and

so that
ax = aH-l(a) x f-l(b) = {O} x f-l(a) x f-l(b).

Consequently, ~18X == cPa,b, and from deg(~18x) = 0 we deduce

(3) We have fixed the center of projection at the south pole and need no
more freedom for our purposes, but this choice is actually irrelevant. The
122 III. The Brouwer-Kronecker degree

argument to show this is quite similar to those above, using a rotation to


move the center elsewhere. We will not give any details here.
(4) In odd dimension, the Hopf invariant vanishes: if m is odd, H (J) =
o for every smooth mapping f : §2m-l -+ §m.
Indeed, we have the commutative diagram

f-l(a) x f-l(b) ~ §2m-2

T 1~ ~ lu
f-l(b) X f-l(a) ~ §2m-2

where T is the permutation (x, y) t---+ (y, x) and u is the antipodal mapping
z t---+ -z. But an easy computation shows that deg(T) = (_1)(m-l)2 = +1
(m is odd) and deg(u) = -1 (111.1.6(1), p. 99), so that
H(J) = deg(<pa,b) = deg(u) deg(<Pb,a) deg(T) = - deg(<pb,a) = -H(J),

and hence H (J) = O. o

Next we see that we have constructed a homotopy invariant:


Proposition 5.3 (Homotopy invariance). Let f, g : §2m-l -+ §m be two
homotopic smooth mappings. Then, H(J) = H(g).

Proof. Since [0,1] X §2m-l is compact, the homotopy (which we can take
to be smooth) is uniformly continuous; hence there is an c > 0 such that

IIHt(x) - Hs(x) II <! for all x E §2m-l and It - 81 <c.


The result follows immediately if we see that for 0 < 8 - t < c, H t and Hs
have the same Hopf invariant. In other words, we can merely assume

IIHt(x) - Hs(x) I <! for all x, t,8.

Now, the result is trivial if H(J) = H(g) = 0; hence we suppose that


one of the mappings is surjective, say f. We will use a boundary argument
(as in the proof of 111.5.1, p. 119) to deduce first

and then
5. The Hopi invariant 123

Notice that what matters here is only that the smooth mappings
~.X = H-l(a) x f-l(b) ---t §2m-2 . (t X y) H ¢(y)-¢(x) and
. . " 1I¢(y)-¢(x) II

are well defined. But by the bound that we have assumed, this follows
immediately if the regular values verify Iia - bll ~ and are taken off !
H ([0, 1] x {-p}) (and the latter is possible by the Easy Sard Theorem
(11.3.4, p. 63), because m ~ 2).
Apart from this, we only recall what was stated in 111.5.2(1), p. 121,
for non-surjective mappings. As mentioned there, if 9 is not surjective, we
can take the regular value a off g(§2m-l), so that 8H-l(a) = {O} x f-l(a);
hence in the first step we get

and this case is finished here. o


As for degree, the preceding homotopy invariance result gives grounds
for defining the Hopf invariant of arbitrary continuous mappings.

Exercises and problems


Number 1. Let f : §2m-l --+ §m be a smooth mapping (m ;::: 2). We assume both
spheres oriented. Let w be a form on §m of maximum degree m, such that w = l. Ism
Then j*w is a closed form in §2m-l, hence exact (because Hm(§2m-l, R) = 0), and there
is a form a of degree m - 1 such that da = j*w. Consequently, we have on §2m-l the
2m - 1 form a 1\ da and can compute the integral

g(f) = 1
s2m-l
a 1\ da.

Show that g(f) is the same for any choice of w and a under the conditions above. How
does it depend on the orientations?
This is in fact another definition of the Hopf invariant, but we cannot prove this now
and consequently we use a different notation.
Number 2. Show that the invariant g defined in the preceding problem is a homotopy
invariant and thus it can be defined for arbitrary continuous mappings.
Number 3. Compute the invariant g(f) of a continuous mapping f : §2m-l --+ §m when
m is odd.
Number 4. Compute H(f) and (!(f) for the mapping f : §3 --+ §2 defined by
f(Xl,X2,X3,X4) = (2Xl\!1- x~ - x~, 2x2V1- x~ - x~, 1- 2x~ - 2x~).
124 III. The Brouwer-Kronecker degree

6. The Jordan Separation Theorem


We devote this section to one very important application of degree theory:
a proof of the Jordan Separation Theorem for closed differentiable hyper-
surfaces of the Euclidean space. In fact, this is the proof involving the
Cauchy index that triggered Hadamard's, and then Brouwer's, research on
the notion of degree (1.2, p. 14). In fact, this proof does not require the no-
tion of degree in full, but only a simpler version which we quickly describe
next.

(6.1) Mod 2 degree theory. All constructions and definitions in the pre-
ceding sections can be carried over disregarding orientations, which means
that manifolds need not be orientable. The resut is a mod 2 invariant de-
fined as follows:
Let f : M ---t N be a proper mapping of two boundaryless differen-
tiable manifolds of dimension m; furthermore, N is connected. Let
g : M ---t N be a proper differentiable mapping close to f and let
a E N be a regular value of g. Then the parity

depends only on the homotopy class of f and is called the degree


mod 2 of f.
We leave the reader the task of reviewing all proofs involved and only
notice for the key result, 111.1.2, p. 96, that the curves r used there have
either no boundary points or two boundary points; hence their union C has
an even number of boundary points.

Using this mod 2 degree, we can now prove the following:


Theorem 6.2 (Jordan Separation Theorem). Let M c jRm+l be a com-
pact, boundaryless, differentiable hypersurface. Then M disconnects jRm+l .

Proof. Pick any point p ¢. M and define the differentiable mapping


x-p
fp : M ---t §m : x t---+ Ilx _ PI! '
which is proper because M is compact. The mod 2 degree of this mapping is
called the mod 2 winding number of M around p and is denoted by w2(M,p)
(compare IV.4.1, p.156). It is easy to see that the winding number is
constant on each connected component of jRm+ 1 \ M.
6. The Jordan Separation Theorem 125

Indeed, if p and q are in the same component, that component, which


is an open set in Rm+1, contains a continuous (even piecewise linear) path
, = ,(t) with ,(0) = p and ,(1) = q. Thus we can define a homotopy by
m x-,(t)
H: [0,1] xM -t § : x t-+ Ilx _ ,(t)ll·
Hence, by homotopy invariance for mod 2 degree,

Thus, to show that M disconnects, it is enough to find two points p and q


with different mod 2 winding numbers.
To that end, pick any point a EM and line Rthrough a whose direction
is perpendicular to TaM. Then pick pER \ M. We see that a is a regular
point of JP by direct computation. For v E TpM we have

daf (v) = Iia - pl12v - (a - p,v)(a - p) = v


p lIa - pll3 Iia - pll '
because (a - p, v) = 0 by the choice of R. Thus, daJp is a linear isomor-
phism, as desired.

Consequently, by the Inverse Mapping Theorem, JP is a local diffeomor-


phism at a. This implies that Jp(M) has non-empty interior in §m, and
by the Sard-Brown Theorem, JP has some regular value U E Jp(M). We
denote by L the half-line x = p + AU, A > o. Then, if p is far enough from
our compact hypersurface M, we have J;l(u) = M n L, and
126 III. The Brouwer-Kronecker degree

Now let x be the first point we find in M n L when starting from p, and
pick a second point q E L \ M before the next (we can do this, because
M n L = f;l(u) i= 0). In this situation,

f;;l(u) = M n L \ {x},
and u is a regular value of fq.
Indeed, the first assertion is clear, and for the second note that

y - q = /-l(Y - p) with /-l > 0 for every y E f;;l(u).

Then, for v E TyM we have

d.f, (v) = Ily-qI12v- (y-q,v)(y-q)


y q Ily - qll3
= Ily - pl12v - (y - p, v)(y - p) = Id f: (v)
/-lIly - pl13 tL Y P ,

which shows that dyfq = tdyfp is an isomorphism.


Thus,

Obviously this implies w2(M,p) i= w2(M, q), and the proof is finished. D

As it stands, the above proof cannot be used for non-compact hyper-


surfaces: for them, fp is not proper any more. However, the idea behind it
can be put to work in that case too. In fact, we have the following:

Theorem 6.3 (Jordan Separation Theorem). Let M c jRm+1 be a closed,


boundaryless, differentiable hypersurface. Then M disconnects jRm+l.

Proof. We will use the following notation. Let D c jRm+1 be an open ball
with boundary a sphere S = D \ D, and let p E jRm+1 \ D. We denote by
7rp the conic projection with center p into the sphere S. This mapping is
defined on a compact cone Kp with vertex p and its image is the spherical
region Kp n S. The boundary of Kp is the tangent cone to the sphere with
vertex p, and the interior Tp of Kp in jRm+1 is the open cone generated by
p and the interior Np of Kp n Sin S; note that Np = 7rp(Tp). We need the
following simple property of 7rp:

dx 7rp(u) = 0 if and only if u is parallel to x - p.


6. The Jordan Separation Theorem 127

Indeed, consider the curve germ ,(t) = x + tu. Then

for some smooth >.(t), and

dx 7rp (u) = (7rp 0,)'(0) = >.'(O)(x - p) + >'(O)u = v.


Since 7rp (x) = p + >'(O)(x - p) E Sand p ~ S, we have >'(0) -:j= 0, and v is
vanishes if and only if u is parallel to x - p.
We can reformulate the above remark by saying that d x 7rp is injective
on any linear hyperplane L transversal to x - p.

Now, pick a point e E M and a line e through e not tangent to M


(that is, e ¢. e + TeM). This choice guarantees that e is isolated in en M
(II.2.12, p. 60): there is a segment [a,e] C econtaining e(-:j= a,e), such that
[a, e] n M = {e}. In particular, c is not in the closed set M, and there is
an open ball D centered at e with M n D = 0.

L~~q
[a,q]

Note that for D small enough, 7ro: restricts to a diffeomorphism ho: : Mo: =
M n To: -+ No:. Indeed, for a first choice of D, the preceding remarks on
the derivatives of conic projections tell that deho: is a linear isomorphism.
Consequently, ho: is a local diffeomorphism at the point e, and the claim
follows by shrinking D.
Once this setting is ready, consider an arbitrary point p ~ M u D and
the corresponding 7rp , K p, Tp, Np. If there is some point a E Np such that
[a,p] n M = 0, we write
128 III. The Brouwer-Kronecker degree

Otherwise, Mp = Tp n M =1= 0 is an m-manifold open in M and 7fp induces


a differentiable mapping hp = 7fp IMp : Mp = Tp n M -+ N p. This mapping
is proper (the restriction of 7fp to the compact set Kp n M is proper), and
we write
W2(p) = deg2(hp).
To compute this W2(p), we use regular values. Now, a E Np is a regular
value of hp if and only if h;l(a) is finite and for each x in that inverse
image, the vector a - p is not tangent to M at x. Again this follows from
what was earlier remarked concerning the derivatives of conic projections.
In short we say that [a,p] meets M transversally, and then

W2(p) = #([a,p] n M) mod 2.

Note that there is some overlapping here. Indeed, Mp may be non-empty


and hp will be defined, even if [a, p] n M = 0 for some a E N p. But then a
is a regular value of hp, so that deg2(hp) = o. Of course, the computation
through any other regular value will give the same degree o.
Next, note that our careful choice of data guarantees w2(a) = 1. On
the other hand, take any !3 in the segment (e, 7fa(e)). Clearly, M{3 = 0,
and W2(!3) = o. Thus, the invariant W2 is not constant. We claim that it is
locally constant.
To prove this, fix p. If [a, p] n M = 0 for some a E Np, as [a, p] is
compact and M is closed, we see immediately that [a, q] n M = 0 for q
very close to p. Hence suppose we have the mapping hp well defined, with
a regular value a E Np so that [a,p] meets M transversally. We want the
same for q close enough to p and also that

#([a,q] n M) = #([a,p] n M),

to conclude W2(p) = W2(q). For this we argue as follows. For every x E


[a,p] n M, let "fa,x be the conic projection with center a onto x + TxM: the
same remarks on derivatives of conic projections apply to "fa to deduce that
its restriction to a neighborhood V of x in M is a diffeomorphism from V
onto an open set of x +TxM. This implies that for q close to p, the segment
[a, q] meets M exactly as [a, p] does, and we are done.
Summing up, W2 is locally constant, but not constant, on the open set
jRm+l \ M U D: thus this set cannot be connected. Finally, if the open
set jRm+l \ M were connected, no closed ball would disconnect it, and we
conclude as desired that the complement of the hypersurface M is not
connected. 0
6. The Jordan Separation Theorem 129

From the fact that hypersurfaces disconnect, much more information


can be deduced. We gather all of it into a single statement:

Proposition 6.4. Let M c Rm+! be a connected, closed, boundaryless,


differentiable hypersurface. Then M disconnects Rm+! into two connected
components D and E, whose closures D and E are closed manifolds with
boundary M. Furthermore, M has a global equation f such that

M = {J = O}, D = {J > O}, and E = {J < O}.


In particular, M is orientable and has a unitary global normal vector field

1/ = grad(J) I II grad(J)II·

If M is compact, one of the two connected components is bounded (the


interior of M) and the other is not (the exterior of M).

Proof. As we know (II.2.3, p. 55), M has local equationsj hence every point
in M has an open neighborhood U in R m +! such that Un M = {J = O}
for some differentiable function f : U --+ R with 0 a regular value. In fact,
by II.2.3, p. 55, we can assume that there is a diffeomorphism cp : V --+ U
from an open ball V centered at 0 E Rm +!, such that

cp-l(U n M) = {x E V : Xm+! = O}.


In this situation:

(a) If a differentiable function h : U --+ R vanishes on un M, then hi f


is a well-defined differentiable function on U.

(b) If h : U --+ R is another equation of M in U, then a = hi f has no


zeros in U.

(c) The set U \ M has two connected components, U+ = {J > O} and


U- = {J < O}, so that M n U c U+ n U-.

For (a), we note that h 0 cp(x}, ... , Xm , 0) = OJ hence

(I a
h 0 cp(X) = Jo at (h 0 CP(Xl, ... , Xm, tXm+l)) dt = h*(x)xm+!,

with
h*(x) = 1
1

o
a
a
Xm+l
(h 0 cp(Xl, ... , Xm , txm+d) dt.
130 III. The Brouwer-Kronecker degree

Similarly, f 0 cp( x) = J* (x )xm+1' but in this case we can say more: J* never
vanishes. Indeed, we have

where 7r : IRm+1 -+ IR is the last coordinate projection. If J*(x) = 0, then


f(cp(x)) = 0; hence cp(x) EM and Xm+1 = O. Thus

and cp(x) E Un M would be a critical point of f, a contradiction. Conse-


quently
* h*(x)
h 0 cp(x) = h (x)xm+1 = f*(x) (J 0 cp)(x),

which shows that (hi J) 0 cp = h* I J* is differentiable on V.


For (b), note that if h is an equation itself, the same argument above
shows that h* has no zeros, and neither does h* I J*.
Finally, assertion (c) is evident for the equation h = 7r 0 cp -1 and follows
for f = ah, because a never vanishes, hence has constant sign on the
connected set U.
After these three local properties, we obtain a covering of M by open
sets Ui C IRm+1 with equations Ii as above, to which we add a covering
of IRm +1 \ M by open connected sets with trivial equations = 1. Recall
that by the Jordan Separation Theorem, 111.6.3, p. 126, we know that M
disconnects IRm+1. In fact,

(d) IRm + 1 \ M has exactly two connected components D and E, and

Indeed, we claim first that W \ W = M for every connected component


W of IRm+1 \ M. Note that any other component W' of IRm+1 \ M is open
in IRm +1 and does not meet W; hence it does not meet W either. Thus,
W \ W eM. On the other hand, the open set W cannot be closed in
IRm+1; hence there is some point x E W \ W eM. Let y E M be arbitrary.
As M is connected, there is a chain Uio' ... ,Uir with x E Uio' Y E Ui r , and
Uik n Uik+l n M 1= 0, Uik \ M = Ui~ U Ui~' for every k:
6. The Jordan Separation Theorem 131

From Uio \ M = !
Ui U Ui~ and x E W, we deduce Ui C W or Ui~ C W. !
!
It follows that M n Uio C Ui n Ui~ C W, and looking at a point Xl E
M n Uio n Uill we see Xl E Uh n W; hence Ui~ C W or Ui~ C W. Thus
MnUil C W.
Repeating the argument, we conclude that y E M n Uir C W. This
completes the proof of our claim.
Now, let D = W be a component as in the argument above, and let E
be a second component (recall that M does disconnect Rm+1). Arguing as
!
above, we find Ui C E or Ui~ C E. As D and E are disjoint, there are
two possibilities:

which gives no extra room for anymore components!


Once we have (d), we choose signs Ci as follows:

Case Ui n M = 0: Ci = {+-11 if Ui cD,


if Ui ¢. D (that is, Ui n D = 0).
Case U· nM -I.. 0· Co = { +1 if Ui+ cD,
~ r·. -1 if Ui+ ¢. D (that is, Ui- CD).

With these signs fixed, set 9i = cili, so that

Consequently, the function 9k/9i, which is defined on Uk n Ui and has no


zeros by (a) and (b), is > O.
Finally, let {(}i} be a differentiable partition of unity for the Ui's. The
formula
132 III. The Brouwer-Kronecker degree

gives the equation f : ~m+1 -t ~ that we sought.

First note that the sum is locally finite, and for x E Ui and Ok(X) =f. 0,
it is x E Uk n Ui , and in that intersection 9k/9i > 0; hence the logarithm
exists. Moreover, in every non-empty intersection Ui n Uj we have

By all of this, (*) is a consistent definition.


Now, on Ui we have f = edi' where ei = Ei exp[· .. J has no zero. Thus,
{f = O} n Ui = {Ii = O} n Ui = Ui n M,
and for every x E Ui n M, derivation gives

Summing up, f is a global equation for M. Furthermore, as exp > 0,


we have D = {f > O} and E = {f < O}, and by (d),

{f > O} \ {f > O} = M = {f = O},


that is, {f > O} = {f ~ O}. This is a manifold with boundary {f = O} = M
(II.3.2, p. 62). The same works for {f < O}.
The other assertions in the statement come from II.7.5(1), p. 88. 0

Exercises and problems

Number 1. Complete the details for a mod 2 degree theory as suggested in 111.6.1. Note
that for oriented manifolds deg 2 = deg mod 2.
Number 2. Let. denote a (bilinear) multiplication in ]Rm+l such that x. y = 0 only if
x = 0 or y = O. Fix any nonzero u E ]Rm+l and define the mapping

f : §m --+ §m : x t-+ f(x) = 1\:: :1\.


Show that f is a homeomorphism whose homotopy class does not depend on u. Deduce
that f is homotopic to - f and that m is odd.
Number 3. Let. denote a multiplication in ]Rm+l as above that furthermore is com-
mutative. Consider
f : lH'm --+ §= : x t-+ f(x) = 1\:: :1\.
7. The Brouwer Theorems 133

Show that:
(1) f is a well-defined injective continuous mapping.
(2) deg 2 (f) = 1; hence f is surjective.
Deduce that U m is homeomorphic to §m. As is well known, this is the case only
for m = 1; hence for m ~ 2 such a multiplication does not exist.
Number 4. Let M c IR m +1 be a connected, compact, boundaryless, differentiable hy-
persurface. Denote by D (resp., E) the bounded (resp., unbounded) connected compo-
nent of IR m +1 \ M. For every point p ¢ M set
m x-P
fp : M -t § : x >-t IIx _ pil .
Prove that deg(fp) is 0 if pEE and ±1 for p E D, the sign depending solely on the
orientation of M.
Number 5. Let N C IRm +1 be a closed smooth manifold of dimension n, N =I IRm+1.
Show that:
(1) If n < m, the complement IRm +1 \ N is connected.
(2) If n = m + 1, N has a boundary M = aN =I 0, and if M is connected, so is
IR m +1 \ N. What if M is not connected?
Number 6. Let M c IR m + 1 be a closed, boundaryless, differentiable hypersurface. Sup-
pose that M has r connected components. How many does its complement IR m +1 \ M
have?

Number 7. Let M c IR m +1 be a connected, boundaryless, smooth hypersurface. Let


U c IR m +1 be a tubular neighborhood associated to vM, and let p : U -t M be the
corresponding retraction. Let v : M -t IR m +1 be a unitary global normal vector field.
Prove that there is a smooth function 8 : U -t IR such that
8(x)v(p(x)) =x- p(x).
Check that 8 is an equation of Min U and that grad(8) = v.

7. The Brouwer Theorems


We will prove here two famous theorems due to Brouwer. We start with
the following immediate consequence of the Boundary Theorem:
Proposition 7.1. Let M be an oriented differentiable manifold with bound-
ary aM = N. Then N is not a proper retraction of M; that is, there is no
proper mapping p : M -+ N whose restriction to N is the identity.

Proof. Suppose there is such a retraction p : M -+ N. Let N' be a con-


nected component of N, and let M' be the connected component of M that
contains N'. Since N' = p(N') c p(M') c Nand p(M') is connected, we
conclude
N' = p(M') = M' nN = aM',
134 III. The Brouwer-Kronecker degree

because M' is open in M. Furthermore, plM' is proper, because M' is closed


in M. Now apply the Boundary Theorem, III. 1.8, p. 101, with X = M',
Y = 8M', and H = pIM', We get 0 = deg(Hly) = deg(pIN') = deg(Id N,),
which is impossible. 0

The condition that the retraction is proper is essential: for instance, look
at the (non-proper) continuous retraction §l x [0, +00) -+ §l : (x, t) f----t x.
From the preceding result we deduce (see 1.2, p. 20 and p. 28):

Proposition 7.2 (Brouwer Fixed Point Theorem). Let M c jRm+l be


homeomorphic to a closed ball. Then every continuous mapping f : M -+
M has some fixed point.

Proof. Clearly, it is enough to prove the statement for M = {x E jRm+l :


Ilxll ::::; I}, which is a compact manifold with boundary 8M = §m. Suppose
there is a continuous mapping f : M -+ M without fixed points. Then
define a mapping p : M -+ §m as shown in the picture

p(x) = f(x) + A(X)(X - f(x))

with '\(x) > 0 such that IIp(x)11 = 1. This mapping is proper and fixes
every point in §m. Thus §m is a proper retraction of M, contrary to the
preceding proposition. This contradiction means that f must have some
fixed point. 0

This theorem will be refined in the next chapter (IV.2.9, p. 152) using
the Euclidean degree. For the moment, here there is a nice application of
it:

Proposition 7.3 (Perron-Fr6benius Theorem). If a regular real matrix has


non-negative entries, then it has some positive eigenvalue, corresponding to
an eigenvector whose coordinates are all ~ O.
7. The Brouwer Theorems 135

Proof. Suppose A is an (m + 1) x (m + 1) matrix. Since A is regular, we


can define the following continuous mapping:

h§m §m A·x
: ---t : x HilA. xII .
Now the hypothesis on the entries of A guarantees that h keeps invariant
the quadrant
Q = {x E §m : Xl ~ 0, ... , Xm+l ~ O}.
But this quadrant is homeomorphic to a closed ball, and by the Brouwer
Fixed Point Theorem, it must contain some fixed point x of h. Conse-
quently
A . x = AX, where A = IIA . xii> O.
Thus, A is the eigenvalue and x is the eigenvector we sought. o
We finish the section with the important fact, also proved by Brouwer,
that only on spheres of odd dimension can we find tangent vector fields
without zeros (1.2, p. 20 and p. 28):
Proposition 7.4 (Hedge-hog Theorem). A sphere §m C ]Rm+1 has a con-
tinuous tangent vector field without zeros if and only if m is odd.

Tangent vector field on §2 vanishing at both poles


~=(-y,x,O).

Proof. If m is odd, then m +1 = 2k, and we have the smooth tangent


vector field
ex = (-X2' Xl,··" -X2k, X2k-l) E Tx§m.

Next, suppose that x H ex E Tx§m is a continuous tangent vector field


on §m without zeros. After dividing by its norm, we have Ilexll = 1 for all
x E §m; in other words, we have a continuous mapping §m ---t §m such that
(x, ex) = 0 for all x E §m. By this condition,
136 III. The Brouwer-Kronecker degree

is a well-defined homotopy, and deg(Ho) = deg(Hl)' Clearly, Ho is the


identity and HI the antipodal diffeomorphism; hence we conclude that the
latter has the same degree as the former, that is, +1. But in III.1.6(1), p.
99, we saw this is possible only for m odd. This finishes the proof. 0

Exercises and problems

Number 1. Let M be an oriented differentiable manifold with boundary 8M = N.


Prove that there is no proper mapping p : M -+ N whose restriction to N is homotopic
to the identity.
Number 2. Let D denote an open ball in IRP, D its closure, and S its boundary. Produce
the following examples:
(1) a continuous mapping D -+ D with no fixed point,
(2) a continuous mapping D -+ D with no fixed point in D,
(3) a continuous mapping D x S -+ D x S with no fixed point. Here, where does the
proof of III. 7.2, p. 134, fail?
Number 3. Let K C IR m +1 be a convex compact set. Prove that every continuous
mapping f : K -+ K has some fixed point. Can convex be replaced by connectecl!
Number 4. Let D C IRm +1 be the unit open ball, and let f : D -+ IR m +1 be a continuous
mapping such that (x, f(x)) < 1 for all x E §m. Show that f has some fixed point a E D.
Number 5. Let D c IRm +1 be the unit open ball, and let f : D -+ IRm +1 be a continuous
mapping such that (x, f (x)) < 0 for all x E §m. Show that f has some zero a ED.
Chapter IV

Degree theory in Euclidean spaces


In this chapter we develop Euclidean degree theory. This could be done in full
generality for proper mappings on closures of arbitrary open sets. However, here we
have chosen the simplest way, enough for the applications, and restricted ourselves
to bounded open sets. The reader will find in §§1-3 that everything can be done
with a little calculus (and some very ingenious arguments); on the other hand,
it is quite straightforward to extend the construction to the general unbounded
case, a profitable exercise we recommend to the reader. Along the way we include
Gauss's proof of the Fundamental Theorem of Algebm and a refinement of the
Brouwer Fixed Point Theorem. Next, in §4, we come back to the origins and
define the winding number, which is the agent behind the scenes for the equivalence
of the Brouwer-Kronecker and Euclidean degrees. In §5, we deduce the famous
Borsuk- Ulam Theorem and an important consequence: the Invariance of Domain
Theorem. After this, §6 contains one fundamental computation: the Multiplication
Formula for the Euclidean degree of a composite mapping. Finally, that formula
is used in §7 to deduce a most general purely topological version of the Jordan
Sepamtion Theorem and from it the Invariance of Domain Theorem once again.

1. The degree of a smooth mapping


From now on, D stands for a bounded open set in lRm+1 , and X = D \ D
stands for its (topological) boundary. Our purpose here is to define the
degree of a continuous mapping D --+ lRm+1. This we will do by the usual
method: first we consider smooth mappings, and then we approximate by
them any given continuous mapping. Here, the reader will recognize some
familiar arguments, which are repeated to make the construction fully self-
contained.
Note that, D being compact, a continuous mapping f : D --+ lRm +1 is
proper (hence closed), and in particular f(D) and f(X) are closed in lRm+1.
Furthermore, the norm

Ilfll = max{IIf(x) II : x E D}
-137
138 IV. Degree theory in Euclidean spaces

is well defined. As usual, this norm measures the approximation to func-


tions on D.
As mentioned above, we will consider the case when f is smooth. In
particular we will be dealing with the set R flD C ]Rm+l of regular values
of fiD. Recall that a E RflD when the derivative dxf is bijective at every
point xED with f(x) = a (trivially true if a tJ. f(D)). The Sard-Brown
Theorem, 11.3.4, p. 63, says that RflD is a residual set, which alows us to
find regular values everywhere. On the other hand, although the set C flD
of all critical points of flD (the points at which the Jacobian determinant
vanishes) is closed in D, the set RflD = ]Rm+l \ f(CfI D) need not be open
in ]Rm+l. However we can say something useful: C flD U X is closed in D;
hence f(Cfl D U X) is closed in ]Rm+l (f is proper), and

RflD \ f(X) = ]Rm+l \ f(Cfl D U X)

These remarks will often be implicit in our arguments. Now, down to


the matter of discussion, we introduce the degree of a smooth mapping:

Proposition and Definition 1.1. Let f : D -+ ]Rm+l be a smooth map-


ping, and let a E ]Rm+l \ f(X) be a regular value of fiD. Then f-l(a) is a
finite set (possibly empty), and we define the degree of f by

d(f, D, a) = L signx(f)
xEf-l(a)

(of course, the sum is 0 in case f-l(a) = 0), where

signx(f) = signdet (;:i, J


(x)) = ±1

is the sign of f at X.

Proof. Note that f-l(a) C D, and the derivative dxf is a linear isomor-
phism at every xED with f(x) = a. Thus, by the Inverse Mapping
Theorem, all points of f-l(a) are isolated, and f-l(a) is discrete. But D is
compact, and f-l(a) does not meet X = D \ D; hence f-l(a) is finite. 0

We stress that signA!) tells whether the derivative dxf : ]Rm+l -+ ]Rm+l
preserves orientation (sign = +1) or reverses it (sign = -1) as seen when
we discussed orientations in the general setting of manifolds (11.7.2). Note
also that from this definition it is clear that:
1. The degree of a smooth mapping 139

(1) If flD is a diffeomorphism, then d(f, D, a) = ±1.


(2) If r : jRm+1.-t jRm+1 is a translation, then d(f, D, a) = d(rof, D, r(a)).

Now, we prove a first property of this degree.


Proposition 1.2 (Degree is locally constant). There exists an open neigh-
borhood of a, W C RflD \ f(X), such that d(f, D, a) = d(f, D, b) for all
bEW.

Proof. As remarked above, D is closed in jRm+1, and jRm+1 \ f(D) is open


in jRm+1 and is contained in RflD \ f(X). Now, if f-l(a) = 0, then a E
jRm+1 \ f(D), and there is an open neighborhood W C RflD \ f(X) of a
that does not meet f(D). Consequently, d(f, D, b) = 0 for all b E W.
After this, we can suppose f- l (a) = {Xl, ... , Xr }. Then, by the Inverse
Mapping Theorem, we find a connected open neigborhood V C jRm+1 \f(X)
of a and r disjoint open neighborhoods UI, .. . ,Ur of Xl .... , Xr, respectively,
such that by restriction we have r diffeomorphisms

whose sign is constant (because Uk is connected) and is denoted by Uk. We


claim that
W = V\f(D\UUk)
k
is the open neighborhood we seek. Indeed, we have:

1. W is open, because D \ Uk Uk is a compact set; hence its image is


closed.

2. a E W, because all its preimages Xk are in Uk Uk'


3. Every b E W has exactly r preimages Yk E Uk. Indeed, b has at
least these r preimages, and if it had others, say Y ¢ Uk Uk, then
Y E D \ Uk Uk, so that f(y) E f(D \ Uk Uk) and b = f(y) ¢ W.

Thus, for every b E W we deduce as claimed

d(f,D,b) = Lsignyk(f) = LUk = L signxk (f) = d(f,D,a).


k k k 0

Next, we turn to homotopy. We start with a somewhat technical state-


ment:
140 IV. Degree theory in Euclidean spaces

Proposition 1.3 (Homotopy invariance, 1st). Let H : [0,1] x D ---+ IRm +1


be a smooth mapping, and consider a point a E IRm +1 \ H([O, 1] x X) that
is a regular value of the three restrictions HolD, HIID, and HI[o,ljxD' Then

d(Ho, D, a) = d(HI, D, a).

Proof. Denote as usual Ht (x) = H (t, x) and let

H(t, x) = (CPl(t, x), ... , CPm+l(t, x)).

Since a is a regular value of H!rO,ljxD and C = H-l(a) does not meet X,


this inverse image

is, by 11.3.2, p. 62, a smooth compact curve possibly with boundary

oC = (C n ({O} x D)) U (C n ({1} x D)) == Hi) 1 (a) U HII (a).

rr
We denote by H, ... , the connected components of C. By the classifica-
tion theorem for smooth curves, each = r rk
is either a compact interval
or a circle. In any case, 11.3.3, p. 63, gives a smooth parametrization of r:
'Y: s r-t (t(S),Xl(S), ... ,Xm+1(s)) = (t(s),x(s)), 0 ~ s ~ 1.

In particular, :: never vanishes. Moreover, from the identity CPi(r(S)) == ai


it follows that

(1) dt ~ 0CPi dx j
O -= -d (CPi°'Y ) (s ) -_ -0CPi -+ L..J - - .
ds at ds . OXj ds
J

Now consider the determinant


dt ~ dXm+l
ds ds CIS

.1(s) =
0e.!.
at
0e.!.
{JXl ~
Xm+l

{J<PFt+ l {J<Pm+l {J<Pm+l


----ax;- (JXm+l

We claim that .1( s) never vanishes. Indeed, the last m + 1 rows are in-
dependent because a is a regular value of H, and by (1) the first row is
1. The degree of a smooth mapping 141

perpendicular to the others, hence independent. Thus we conclude that


.1(s) has no zero; hence it has constant sign:
(2) .1(s) has the same sign at s = 0 and s = 1.
Next, we look again at the equations in (1), which tell how to combine
the columns in our determinant to get

~
112112 ds
dxm±l
ds

0fl ~
11~;112 det (:~;b(S))).
dt ..1 = 0 8Xl 8 Xm±1
(3) =
ds
8<pmtl 8<pm±1
0 8Xl 8Xm ±1

After this preparation, we compute degrees:

d(HO, D, a) = ~k ~(O,Y)Erk signy(Ho),


(4) {
d(Hl' D, a) = ~k ~(1,z)Erk signAHl).

For that computation, we are only interested in the components r = rk


that meet the levels t = 0, 1. Those components are the bordered com-
ponents among the rk'S, that is, the components diffeomorphic to a com-
pact interval; hence they have two distinct boundary points, ,,(0) = p and
,,(I) = q, both at those t-levels. The following picture shows the various
possibilities:

Clearly, the required equality of degrees d(Ho, D, a) = d(Hl, D, a) will


come from the following facts:
142 IV. Degree theory in Euclidean spaces

1. If p and q are at the same t-Ievel, they contribute to the same degree
in (4) with opposite signs; hence they cancel each other.
2. If p and q are at different t-Ievels, they contribute to different degrees
in (4) with the same signs in both.

This is easily proved. Consider the first component t : [0,1] -+ [0,1] of


the parametrization 'Y = (t,x). Since (~("((s))) is the Jacobian matrix of
J
Ht(s) and a is a regular value of Ho and HI, the right-hand side in (3) is
not zero for t(s) = 0 and 1; hence ~!(O) i= 0 and ~!(1) i= o. We have:

l(i) If p and q are at the same level t = 0, that is, t(O) = t(l) = 0, then
t is increasing at 0 and decreasing at 1, so ~!(O) > 0 and ~!(1) < o.
From (2) and (3) we deduce that the signs of Ho at x(O) and x(l) are
opposite.
l(ii) If p and q are at the same level t = 1, that is, t(O) = t(l) = 1, then
t is decreasing at 0 and increasing at 1, so ~! (0) < 0 and ~! (1) > O.
By (2) and (3) again, the signs of HI at x(O) and x(l) are opposite.
2(i) If p and q are at levels t = 0 and t = 1, that is, t(O) = 0 and t(l) = 1,
then t is increasing at 0 and 1, so ~! (0) > 0 and ~! (1) > O. By (2)
and (3) as before, we see that the signs of Ho at x(O) and HI at x(l)
are the same.
2(ii) If p and q are at levels t = 1 and t = 0, that is, t(O) = 1 and t(l) = 0,
then t is decreasing at 0 and 1, so ~! (0) < 0 and ~! (1) < o. Thus,
from (2) and (3) we deduce that the signs of Ho at x(l) and HI at
x(O) are the same.

This completes the proof. D

The previous statement can be simplified a bit:


Proposition 1.4 (Homotopy invariance, 2nd). Let H : [0,1] x D -+ ~m+1
be a smooth mapping, and consider a point a E ~m+1 \ H([O, 1] x X) that
is a regular value of the two restrictions HolD and HIID. Then
d(Ho, D, a) = d(H I , D, a).

Proof. This follows from the preceding results. By IV.1.2, p. 139, there is an
open neighborhood W of a that does not meet the compact set H([O, 1] xX)
and consists solely of regular values of HolD and HIID and is such that
d(Ho, D, a) = d(Ho, D, b) and d(H I , D, a) = d(H I , D, b)
1. The degree of a smooth mapping 143

for all b E W. Now, by the Sard-Brown Theorem, W contains some reg-


ular value b of the mapping HI[o,ljXD, and applying IV.1.3, p. 140, we get
d(Ho, D, b) = d(Hl' D, b), which jointly with the preceding equalities proves
the statement. 0

Once this is established, we can improve on IV.1.2, p. 139, and define


degree for non-regular values:

Proposition and Definition 1.5. Let f : D -+ jRm+1 be a smooth map-


ping, and let a E jRm+l \ f(X). Let fl be the connected component of
jRm+1 \ f(X) that contains a. Then d(J, D, b) is constant for all b E
fl n R(JID), and we define the degree of f by

d(J, D, a) = d(J, D, b)

for any such b {which exists by the Sard-Brown Theorem}.

Proof. Let b, b' E fl n R(JID)' Since fl is an open connected set in a


Euclidean space, there is a smooth arc "( : [0,1] -+ fl with "((0) = b, "((1) =
b', and we define

H: [0,1] x D -+ jRm+I : (t, x) M f(x) - "((t).

We have:

(1) 0 rt. H([O, l] x X): if f(x) = "((t), for some x E X, then f(x) E
f(X) n fl, which is a contradiction.

(2) 0 is a regular value of HolD: since Ho = f - b, the derivatives dxHo


and dxf coincide, and we know that b is a regular value of f.

(3) 0 is a regular value of HIID: same argument for HI = f - b'.

Thus, by IV.1.4, p. 142,

d(Ho, D, 0) = d(Hl' D, 0).


Finally, since both Ho and HI are translations of f, we obtain

d(Ho, D, 0) = d(J, D, b) and d(HI' D, 0) = d(J, D, b').


All together, we get what we want. o
For this extended definition, IV.1.4, p. 142, remains true:
144 IV. Degree theory in Euclidean spaces

Proposition 1.6 (Homotopy invariance, 3rd). Let H : [0,1] x D -+ jRm+l


be a smooth mapping, and consider a point a E jRm+ 1 \ H ([0, 1] x X). Then

d(Ho, D, a) = d(HI, D, a).

Proof. Let Q be the connected component of jRm+l \ H([O, 1] x X) that


contains a. By the Sard-Brown Theorem, Q contains some regular value b
of both HolD and HIID, and then

d(Ho, D, a) = d(Ho, D, b) and d(Hl' D, a) = d(Hl' D, b)

(IV.1.5, p. 143). But by IV.1.4, p. 142, d(Ho, D, b) = d(HI, D, b), and we


are done. D

Exercises and problems

Number 1. Consider the quadratic transform

and the open set D c IR n defined by 01 < Xl < (31, ... ,On < Xn < (3n. Describe the function
a r-+ d(f, D, a).
Number 2. ([Picard 1891]; 1.1, p. 12) Let D c IR m +1 be a bounded open set, let X =
D \ D, and let/1, ... ,fm+! : D --+ IR be smooth functions. Suppose that 0 ¢ f(X) is a
regular value of flD, where f = (/1, ... ,!m+1) : I5 --+ IRm+l. Consider the bounded open
set E = D x (-1,1) C IR m+ 1 x IR = IR m+2 and the smooth function

-
fm+2 : E --+ 1R: (x, t) r-+ t . det (Bh
Bx (x) ).
J

Finally, consider

Prove that d(g, E, 0) is well defined and that it coincides with the number of solutions in
D of the system
/1 (x) = ... = fm+l(X) = o.

Number 3. ([Heinz 1959]; lA, p. 38) Let D C IR m +1 be a bounded open set, let X =
D \ D, and let f : D --+ IR m+! be a smooth mapping. Let a E IR m+ 1 \ f(X) be a
regular value of f. Let 'Pe : IR m+ 1 --+ IR be a smooth mapping whose closed support
Ke is contained in the open ball of radius c > 0 centered at the origin and such that
IRm+l 'Pe = 1 (called a convolution kernen. Prove the Heinz Integral Formula:

d(f, D, a) =
JDr 'Pe(f(X) - a) det (~(x»dx
3

for c small enough.


2. The degree of a continuous mapping 145

Number 4. Explain and prove the following product formula for smooth mappings:

d(f x g, D x E, (a, b)) = d(f, D, a) . d(g, E, b).

Number 5. Let D c ll~? denote the open disc of radius 2, and consider the smooth
mapping j = (II, h) : D -+]R2 given by

4 + y)e XY + x 2y - x 2 + y + 1,
{ II (x, y) = (_x
12 (x, y) = x y - 4xy.
3

Study the variation of d(f, D, (a, 0)) for a E ]R, when defined.
Number 6. Let X C ]R2 be the curve union of the following pieces:
(i) the graph of the topological sinus y = sin(l/x), 0 < x :::; y'3,
(ii) the segment x = 0, -1 :::; Y :::; 1,
(iii) the arc of the circle x 2 + (y + 1)2 = 4 off the first quadrant, and
(iv) the segment x = y'3, 0:::; y :::; sin(l/y'3).
Show that X is not homeomorphic to §1, but nonetheless it bounds an open set. Then
prove that the mappings j, 9 : X -+ ]R2 \ {(O, O)} given by

j(x, y) = (x 3 + y3 + 1, x - y) and g(x, y) = (x 3 + y3 + 1, x + y)


are not homotopic.
Number 7. Let D C IR m +1 be a bounded open set and let X be its boundary. Suppose
we are given two smooth functions j, 9 : D -+ IR m +1, a homotopy H : [0,1] x D -+ IRm +1
with Ho = j, HI = g, and a point a E ]Rm+1 \ H([O, 1] x X) that is a regular value of j
and g. Prove that the fibers j-I(a) and g-l(a) have the same parity. Is it always true
that they have the same number of points?

2. The degree of a continuous mapping


After the discussion of smooth mappings in the preceding section, we are
ready to define degree for arbitrary continuous mappings:

Proposition and Definition 2.1. Let f : D --+ jRm+1 be a continuous


mapping and let a E jRm+l \ f(X). Then there exists a smooth mapping
g: D --+ jRm+l such that IIf -gil < dist(a, f(X)). For all such 9 's the degree
d(g, D, a) is defined (a E jRm+1 \ g(X)) and is the same, and we define the
degree of f by
d(f, D, a) = d(g, D, a).
Furthermore, 9 can be chosen such that a is a regular value of giD' and
then
d(f, D, a) = signx(g)·L
xEg-I(a)
146 IV. Degree theory in Euclidean spaces

Proof. First of all, by the Weierstrass Approximation Theorem, we find a


polynomial (hence smooth) mapping P such that

liP - fll = max {IIP(x) - f(x)11 : x E D} < ~ dist(a, f(X))

(recall that a ¢. f(X) and f(X) is closed). Next, by the Sard-Brown


Theorem, there is some regular value b of PID such that

Iia - bll < ! dist(a, f(X)).


One easily checks that 9 = P +a - b is the function we sought. To see that
d(g,D,a) = d(g',D,a) for two mappings g,g' that are close to f, consider
the smooth homotopy

H: [a, 1] x D --+ ]Rm+1 : (x, t) H (1 - t)g(x) + tg'(x).


Clearly, Ho = 9 and HI = g'; hence to apply IV.1.6, p. 144, we only must
check that a ¢. H([a, l] x X). But suppose there is an x E X such that
a = Ht(x). Then

Iia - f(x)11 = IIHt(x) - ((1- t)f(x) + tf(x)) II


:S (1- t)llg(x) - f(x)11 + tllg'(x) - f(x)11
< (1 - t) dist(a, f(X)) + tdist(a, f(X)) = dist(a, f(X)),
a contradiction. D

Remark 2.2. As is to be expected, the approximations 9 used above to


define the degree of f are homotopic to f off a on X. Indeed, the homotopy

Ht(x) = tf(x) + (1 - t)g(x)


does not touch a. For, if there is an x E X with Ht(x) = a, then a lies in
the segment between f(x) and g(x); hence

Ilf(x) - g(x)11 2 Ilf(x) - all 2 dist(a, f(X)),

contrary to the choice of g. D

From the definition, we get the following:

Proposition 2.3. Let f : D --+ ]Rm+1 be a continuous mapping. Then the


degree d(f, D, .) is constant on every connected component of ]Rm+1 \ f (X).
2. The degree of a continuous mapping 147

Proof. Let a E lR,m+l \f(X), and, according to IV.2.1, p. 145, pick a smooth
mapping 9 : D -+ lR,m+1 such that Ilf -gil < dist(a, f(X)). We can compute
the degree d(f, D, a) through g, but in fact, we can also compute d(f, D, b)
for any other bE lR,m+1 \ f(X) close enough to a: namely whenever

Iia - bll < TJ = dist(a, f(X)) - Ilf - gil·

Indeed, this implies

dist(b, f(X)) ~ dist(a, f(X)) - Iia - bll > dist(a, f(X)) - TJ = Ilf - gil,

and since TJ:::; dist(a,g(X)) (check it!), the two points a,b are in the same
connected component of lR,m+l \ g(X). Then, by IV.1.5, p. 143, we see that

d(f, D, a) = d(g, D, a) = d(g, D, b) = d(f, D, b).

This means that the function a r--+ d(f, D, a) is locally constant on lR,m+l \
f(X), and, since its values are integers, it must be constant on every con-
nected component of lR,m+1 \ f(X). 0

Of course, we can improve on the homotopy invariance once the defini-


tion is most general:

Proposition 2.4 (Homotopy invariance). Let H : [0,1] x D -+ lR,m+l be a


continuous mapping, and let 'Y : [0, 1] -+ lR,m+1 be a continuous path such
that 'Y( t) ¢ Ht (X) for 0 :::; t :::; 1. Then d( Ht, D, 'Y( t)) does not depend on
t.

Proof. Fix 0:::; to :::; 1. The function (t,x) r--+ Ib(t) - Ht(x)11 is continuous
and> 0 on the compact set [0, 1] x X and hence has a minimum c > O. On
the other hand, H is uniformly continuous on the compact set [0,1] x D;
hence there is an TJ > 0 such that

IIHt(x) - Ht'(X') II < !c if It - til, IIx - xiII < TJ.


Then, by the Weierstrass Approximation Theorem, there is a polynomial
mapping 9 such that

Ilg(x) - Hto(x)11 < !c for x E D.

We deduce that
148 IV. Degree theory in Euclidean spaces

for x E D, It - tol < TJ. Thus,


Ilg(x) - Ht(x) II < dist(r(t) , Ht(X)) for x E D, It - tol < TJ.
In particular, for x EX, t = to, we get

hence g(x) =1= l'(to). As g(X) is compact, we deduce that

dist(r(to), g(X)) > o.


Thus, for a possibly smaller TJ, we can assume that It - tol < TJ implies

Ib(t) -1'(to)11 < dist(r(to), g(X)).


Now we are ready to compute the degree. Suppose It - tol < TJ. First, in
view of the construction, we can apply IV.2.1, p. 145, to get

d(Ht,D,I'(t)) = d(g,D,I'(t)).
Then, by translation

... = d(g -1'(t), D, 0).

Next,
... = d(g - l'(to), D, 0),
because

II(g -1'(t)) - (g -1'(to))11 = Ib(t) -1'(to)11


< dist(r(to),g(X)) = dist (0, (g -1'(to))(X))
and IV.2.1, p. 145, applies again. We continue the computation, and by
translation
... = d(g, D, l'(to)).
Finally, by a third use of IV.2.1, p. 145, we obtain

Hence, d(Ht, D, l'(t)) is locally constant and consequently constant in the


connected interval [0,1]. 0

We have thus defined the degree and proved its homotopy invariance
in full generality. As an easy consequence we deduce the other properties
that characterize degree axiomatically:
2. The degree of a continuous mapping 149

Corollary 2.5 (Nagumo, Fiihrer, Deimling Axiomatizations).


(1) Normality: If f is the inclusion IdD : D -+ D c IRm +1, then

d(f D a) =
"
{I
0
if a E D,
if a fJ. D.

(2) Existence of solutions: Given a continuous mapping f : D -+ IRm +1


and a point a fJ. f(X), if d(f, D, a) =1= 0, the equation f(x) = a has some
solution in D.
(3) Additivity: Given two disjoint open sets Dl, D2 CD, a continuous
mapping f: D -+ IRm +1 , and a point a fJ. f(D \ Dl U D2), then
d(f, D, a) = d(f, D 1 , a) + d(f, D2, a).
Proof. For (1) use IV.I.1, p. 138.
For (2) we argue by way of contradiction. Suppose f-l(a) = 0. Then
dist(a, f(X)) ~ dist(a, f(D)) = E > 0,
and by the Weierstrass Approximation Theorem there is a polynomial map-
ping 9 = g(x) such that Ilg - fll < !E;
in particular, a fJ. g(X). Then,
d(f, D, a) = d(g, D, a). For x E D we have
Iia - g(x)11 ~ lIa - f(x) II -lIg(x) - f(x)11 > E- !E = !E.
This means that dist(a,g(D))> !E.
Now, we pick, in the connected com-
ponent of IR +1 \ g(X) that contains a, a regular value b of glD such that
m

lIa - bll <!E. Then g-l(b) = 0 and d(g, D, a) = d(g, D, b) = 0, a contra-


diction.
For (3) notice first that D \ Dl U D2 contains the three boundaries

D\D, Dl \Dl' D2 \D2.


Now, dist(a, f(D \ Dl U D2)) = E > 0, and we can choose a polynomial
mapping 9 = g(x) with Ilg - fll < E. Then a fJ. g(D \ Dl U D2), and it
follows that
d(f, Dl, a) : d(g, D 1 , a),
d(f, D, a) = d(g, D, a) and {
d(f, D2, a) - d(g, D2, a).

Finally, let [} be the connected component of Jlrl+1 \ g(D \ Dl U D2) that


contains a, and pick in [} a regular value b of giD. Then

d(9' Dl, a) : d(g, D 1, b),


d(g, D, a) = d(g, D, b) and {
d(g, D2, a) - d(g, D2, b).
150 IV. Degree theory in Euclidean spaces

All in all, assertion (3) follows from IV.2.1, p. 145, because

o
We complete the construction of the Euclidean degree with another very
important fact:
Proposition 2.6 (Boundary Theorem). Given two continuous mappings
f,g : D -+ ~m+1 such that fix = glx and given a point art. f(X) = g(X),
then
d(f,D,a) = d(g,D,a).

Proof. Apply IV.2.4, p. 147, to the homotopy Ht(x) = tf(x) + (1 - t)g(x)


and the path 'Y == a. Note that H t = f = 9 on X. 0

Example 2.7 (Gauss). We come back to the solution of complex algebraic


equations, already discussed earlier (III. 1. 10, p. 102). There we used the
Brouwer-Kronecker degree for proper mappings, but we can instead apply
the more elementary Euclidean degree, along Gauss's lines (see 1.1, p. 2).
Let P(z) = zP + CIZp-1 + ... + cp be a polynomial with complex coeffi-
cients. If ( E C is a root of P with 1(1 ~ 1, then

hence
1.,1"1 -< ICIII(IP-l1(lp-1
+ ... + Icpl < Ic 1+··· + Ic I.
- I p

Thus we set p = 1 + ICII + ... + ICpI, and consider in ~2 == C the open


disc D = {Izl < p}, with boundary X = {Izl = pl· Then f = PI D is
a continuous mapping with 0 rt. f(X), and we are to compute the degree
d(P, D, 0). To that end consider the homotopy
Pt(z) = zP + tCIZp-1 + ... + tcp, 0::; t ::; 1.
Using the same bound on the coefficients, we see that Pt has no root in X;
hence (IV.2.4, p. 147)

d(f, D, 0) = d(PI , D, 0) = d(Po, D, 0).


Now the mapping Po : z f---+ zP has degree p, for, any small enough a i= 0
is a regular value of Po with p different roots c in D. Moreover, at each c,
Po preserves orientation: dcPo is multiplication times the complex number
pcp-I.
2. The degree of a continuous mapping 151

We have thus seen that the degree of a complex polynomial is exactly


its ordinary degree as a polynomial. Quite predictable, but very important,
is the fact that degree is not zero; hence f = PID is surjective onto the
connected component [l of C \ f(X) that contains 0 (IV.2.5(2), p. 149). In
particular, P has some root in D: this is again the Fundamental Theorem
of Algebra.
Furthermore, we can even see that degree counts the number of roots
of P. Indeed, the derivative of f = PID at any given point z is the linear
mapping defined by complex multiplication by P'(z); such a multiplication
preserves orientations. Thus, at every regular value a E [l of f we get

p = d(j, D, a) = L signA!) = #f-1(a);


zEf-1(a)

that is, the equation P(z) = a has p roots. If a is not a regular value, the
roots must be counted with multiplicities; details are left to the reader. 0

We close this section with a variation on the Brouwer Fixed Point The-
orem as was stated in 111.7.2, p. 134. First we formulate a somewhat tech-
nical version:

Proposition 2.8. Consider the unit open ball D m +1= {x E lRm +!: Ilxll < I}
and the unit sphere §m c lRm + 1 .

(1) Let f : yyn+l -+ lRm +! be a continuous mapping such that f(x) =F 0


for all x E yyn+ 1 . Then there is a z E §m such that f (z) = AZ for some
A> O.
(2) Let g : yyn+! -+ lRm + 1 be a continuous mapping such that g(x) =F x
for all x E DIn+!. Then there is a z E §m such that g( z) = AZ for some
A> 1.

Proof. (1) Suppose by way of contradiction that f(x) =FAX for all x E §n
and A > O. Then we can apply IV.2.4, p. 147, to the homotopy

Ht(x) = (1 - t)f(x) - tx,

so that

d(j, D m +!, 0) = d(Ho, D m +!, 0) = d(Hl' D m +1, 0)


= d(-Id1r+l,D m +!, 0) = (_I)m+l =F O.
152 IV. Degree theory in Euclidean spaces

Hence, by IV.2.5(2), p. 149, there must be some x E Dm+l with f(x) = 0,


a contradiction.
(2) Apply (1) to the mapping f(x) = g(x) - x. o
Now the mentioned variation on 111.7.2, p. 134, is immediate:
Corollary 2.9. Let D c ffim+l be an open ball with boundary S. Then
every continuous mapping f : D -+ ]Rm+l such that f(S) C D has some
fixed point.

Proof. Let Dm+l be the unit open ball, and consider the mapping

g:
=D+1 TlDm+l
-+ll'Io.: xt-+
f(rx + c) - c
,
r
where c is the center of D and r is its radius. By the hypothesis on f,
g(§m) C D, so that for all x E §m we have IIg(x)II ~ 1; hence g(x) =1= AX if
A> 1. Consequently, by IV.2.8(2) above, 9 must have some fixed point x,
and rx + c is a fixed point of f. 0

Remark 2.10. We have thus weakened the condition f(D) C D to f(S) c


D, and this is indeed weaker. For instance, look at

g(x) = {3X for Ilxll ~ !,


(5 - 411xll)x for Ilxll ~ !.
This mapping fixes §m but sends all points in i < Ilxll < 1 off Ilxll ~ 1 .0

Exercises and problems


Number 1. Extend Euclidean degree to unbounded domains starting with the following.
Let a c Rm + l be open, possibly not bounded. Then consider continuous mappings
f : ?i -+ Rm+l verifying
sup IIx - f(x)11 < +00.
",En

Prove that for every a 1. f(8?i), f-l(a) is bounded and for all bounded open sets Dca
containing f- l (a), the degrees d(J, D, a) are the same. This common value is the degree
d(J, a,a).
Number 2. ([Poincare 1883]; 1.1, p. 11) Let al, ... ,an be positive real numbers, and
for each i = 1, ... , n, let Ii : [-al, all x ... x [-an, an] -+ R be a continuous function
positive on Xi = ai and negative on Xi = -ai. Prove that the system
/l(X) = ... = fn(x) = 0
3. The degree of a differentiable mapping 153

has some solution in (-al,al) x ... x (-an, an).


Number 3. Let DC ]Rm+1 be a bounded open set, let X = D\D, and let J : D --t ]Rm+1
be a continuous function such that J(D) is contained in a hyperplane. Then d(f, D, a) =
for all a E ]Rm+l - J(X).
°
Number 4. Let a < b be two real numbers and let D be a bounded open set of [a, b] x
]Rm+1 that meets all sets {t} x ]Rm+1. Let H : D --t ]Rm+1 be a continuous function and
let c E ]Rm+1 \ H(8D). For a ~ t ~ b and X C [a, b] x ]Rm+1 we denote
X t = {x E ]Rm+l: (t,x) EX}.
Show that D t C ]Rm+l is a bounded open set, that 8D t C (8Dh (equality not true in
general), and that the degree d(Ht, D t , c) is well defined and does not depend on t.
Number 5. Let D be a bounded open subset of [0, 1] x ]Rm+l meeting all sets {t} x ]Rm+1 ,
and let H : D --t ]Rm+1 be a continuous function. Let a E ]Rm+1 off H t (8D t ) for all
t E [0,1]. Is it true that d(Ht, D t , a) depends on t?

Number 6. Write]Rm =]RP x ]Rq and consider the closed balls


D: = {x E ]RP : Jlxll ~ c} and D~ = {y E ]Rq : IlyJl ~ a}.
Let there be a continuous mapping
g: D: x D~ --t]Rm: (x,y) f-t (x,gy(x))
with g(O, y) = y. Show that
Ugraph(gy) = {(x,gy(x)) : x E D:,y E Dn
y

is a neighborhood of (0,0) in ]Rm.


Number 7. Let D denote the unit open ball in ]Rm+1, and let J : D --t ]Rm+l be a
continuous function without fixed points. Prove that the angle a(z) determined by the
°
two vectors z, z - J(z) reaches all values from to 11" for z E §m.

3. The degree of a differentiable mapping


Only after all properties of the Euclidean degree are available, can we de-
duce that for an arbitrary differentiable mapping degree can be computed
by the same regular value formula that gives the degree of a smooth map-
ping. In any case, it is interesting to remark how simple this is if we compare
it to the Brouwer-Kronecker theory (III.4.1, p. 114). We need the following
lemma, interesting in its own right:
Lemma 3.1. Let D c ]Rm+1 be the open ball Ilx - ell < c, and let f :
{l --t ]Rm+1 be a C1 mapping defined on a convex open neighborhood {l of
D. Then the mapping
f(tx + (1 - t)c) - f(c)
for t =1= 0,
Ht(x) = { t
dcf(x - c) for t =0
154 IV. Degree theory in Euclidean spaces

is a continuous homotopy on n of f - f(c) and de! - de!(c).

Proof. Since n is convex, we can write


f(x) - f(c) = 11 !f(tx + (1- t)c)dt

= 1 1 m+1
L - . (tx
o i=1 dx~
of
+ (1 - t)C)(Xi - ci)dt =
m+l
L (Xi - Ci)!i(X),
i=1
where the !i's are the continuous functions

fi(X) = 1o
1 of
-d' (tx
x~
+ (1 - t)c)dt.

We see that fi(C) = ~(c), and after a little computation we get


m+l
Ht(x) = L (Xi - Ci)fi(tx + (1 - t)c),
i=1
which shows that H is continuous. D

Now we can prove the following:


Proposition 3.2. Let f : D -+ IRm +1 be a continuous mapping whose
restriction flD is C1, and let a E IRm +1 \ f(X) be a regular value of fiD.
Then f-l(a) is finite, and

d(J,D,a) = L signdet (;:i. (x)).


XE!-l (a) J

Proof. That f-l(a) is finite follows as usual from the Inverse Function The-
orem, say f- 1 (a) = {Cl, ... , Cr }. We pick disjoint open balls Di centered
at the points Ci with Di c D, such that each restriction flDi is bijective.
Since a tj. f(D \ Ui Di), from additivity (IV.2.5(3), p. 149), we deduce that

d(J,D,a) = Ld(J,Di,a).
i

Then, by the previous lemma and the invariance of homotopy, we have


d(J, D i , a) = d(J - J(Ci), Di, 0)
= d(dcJ - dcJ(Ci), Di, 0) = d(deJ, Di, dcJ(Ci)).
As (~( cd) is the matrix of the linear isomorphism deJ, we get what we
J
wanted. D
3. The degree of a differentiable mapping 155

Exercises and problems

Number 1. Let hI, ... ,hr : JR. -+ JR. be differentiable functions, and define

f : JR.2 -+ JR.2 : (x, Y) H (xy, xy2 + x 2 I>i(xy)2).


Compute all posible values of d(f, D, a).

Number 2. Let f : D -+ JR. m+1 be a continuous mapping and consider a E JR. m+ l \f(X).
Suppose that f is differentiable on an open set containing f- l (a). Prove that

d(f,D,a)= L Signdet(;~i(x))
xEJ-1(b) J

for all regular values b close enough to a.

Number 3. Let f : D -+ JR. m+ l be a continuous mapping whose restriction flD is


differentiable, with Jacobian determinant 2 0 all through D. Let a E JR. m+ l have finite
preimage f-l(a) c D. Is it true that

d(f,D,a) = L signdet (;~i(X))


xEJ-1(a) J

(where sign 0 = O)?


Number 4. Let f : D -+ JR. m+ l be a continuous mapping whose restriction flA to an
open set A c D is differentiable, and let a E JR. m+1 \f(X) be a regular value of fiA. Find
conditions on the fiber f- l (a) for the following formula to hold true:

d(f,D,a) = L signdet (;~i (x)).


xEJ-1(a)nA J

Number 5. Let f : D -+ JR.m+l be a continuous mapping whose restriction flD is


differentiable. Suppose that the open set A consisting of all regular values of f in JR. m+1 \
f(X) is connected. Prove that

for any two values a, b E A.

Number 6. Let Dk C JR.k denote the unit open ball as usual. Let f : I5 P -+ JR.P be a
continuous mapping whose restriction flDp is differentiable with 0 ~ f(f5P\DP) a regular
value. Consider g : D P + q -+ JR.p+q defined by

Prove that d(g, Dp+q, 0) is well defined and that it coincides with d(f, DP, 0).

Number 7. ([Fiihrer 1971]; 1.5, p. 39) We already know that Euclidean degree theory
verifies all properties of Fiihrer's characterization. Prove now that those properties give
uniqueness of the theory, by the following steps:
(1) By approximation and homotopy invariance, reduce uniqueness to the case of
differentiable functions.
156 IV. Degree theory in Euclidean spaces

(2) By additivity, reduce to mappings defined on open balls and regular values with
a single preimage.
(3) By IV.3.1, p. 153, reduce to linear forms, and then use paths in the space of
regular matrices.

4. Winding number
Here we look at the relationship between the Brouwer-Kronecker degree
and the Euclidean degree, where our X in this chapter is a hypersurface.
The basic notion on which the whole comparison stands is that of winding
number, and thus we come back to the origins of the theory (1.1, p. 5, and
1.2, p. 19).
First of all, for D C ~m+1 bounded open and X =D \ D as usual, we
have the following:
Proposition and Definition 4.1. Let f : X -+ ~m+l be a continuous
mapping and consider a point a E ~m+1 \ f(X). Then, for all continuous
extensions J: D -+ ~m+1 the degree d(J, D, a) is the same, and we define
the winding number of f around a by

w(j, a) = d(J, D, a)

for any such J (which exists by the Tietze Extension Theorem).

This follows immediately from the Boundary Theorem, IV.2.6, p. 150.


As is the degree (IV.2.3, p. 146), the winding number is also locally constant
on the target:
Proposition 4.2. Let f : X -+ ~m+l be a continuous mapping and con-
sider two points a, b in the same connected component of ~m+1 \ f(X).
Then w(j, a) = w(j, b).

Of course, the winding number is invariant by homotopy:


Proposition 4.3. Let H : [0,1] x X -+ ~m+1 be a continuous mapping,
and let "( : [0,1] -+ ~m+l be a continuous path such that "((t) Ht(X) fortt
0:::; t:::; 1. Then w(Ht, "((t)) does not depend on t.

Proof. By the Tietze Extension Theorem, H extends to a homotopy H :


[0,1] x D -+ ~m+l, and IV.1.3, p. 140, applies. D

Another easy fact is the following consequence of IV.2.5(2), p. 149:


4. Winding number 157

Proposition 4.4 (Boundary Theorem). Let f : X -t ]Rm+1 be a continuous


mapping and consider a point a E ]Rm+l \ f(X). If f has a continuous
extension f: D -t ]Rm+1 such that a ¢:. f(D), then w(j,a) = O.

Next we want to represent the Brouwer-Kronecker degree as a winding


number in this Euclidean setting. The key result needed to do this follows.
Proposition 4.5. Let X be the boundary of a compact, oriented, differ-
entiable manifold W of dimension m + 1. Let f : W -t ]Rm+1 be a dif-
ferentiable mapping and let a E ]Rm+1 \ f(X) be a regular value of f, and
consider the differentiable mapping
f(x) - a
g : X -t §m : x H .
Ilf(x) - all
Then f-l(a) is finite, and the Brouwer-Kronecker degree of g is

deg(g) = L sign x (/)'


xEf-l(a)

Proof. That f-l(a) is finite follows as usual, say f-l(a) = {Xl, ... .x r }.
Then there are disjoint neighborhoods Uk of the points Xk such that the
restrictions fluk : Uk -t B are diffeomorphisms onto an open ball B cen-
tered at a. Then, we choose a smaller closed ball DeB with boundary a
sphere S and again restrict f to have diffeomorphisms
flvk : Vk = f-I(D) n Uk -t D, with f(aVk) = S.
Let V = W \ Uk Vk. This is a compact manifold with boundary Y =
X U Uk aVk; here X and the aVk'S carry the orientation as boundaries of
V. This means that each aVk carries the wrong orientation as boundary of
Vk·
158 IV. Degree theory in Euclidean spaces

Since a 1. j(V), 9 is in fact defined on V, and by III. 1.8, p. 101,


deg(gI8v) = 0; hence

deg(glx) - Ldeg(gI8Vk ) = 0,
k

where the negative sign corresponds to the above remark on the orientations
of the 8Vk'S. Moreover, if c is the radius of D, we have
1 -
gl8Vk = e(j1 8Vk - a),
and it follows readily that deg(gI 8vk) = deg(J18vk). Now:

(1) If signxk(J) = +1, the diffeomorphism JIVk : Vk -+ D preserves ori-


entation. Hence the restriction JI 8Vk : 8Vk -+ S preserves orientation
too and has Brouwer-Kronecker degree +1.

(2) If signxk(J) = -1, the diffeomorphism JIVk : Vk -+ D reverses ori-


entation. Hence the restriction JI 8Vk : 8Vk -+ S reverses orientation
too and has Brouwer-Kronecker degree -1.

Consequently,

deg(glx) = Ldeg(gI8vk) = Ldeg(j18vk) = Lsignxk(J),


k k k
as desired. o
Now we conclude the section with the announced representation of the
Brouwer-Kronecker degree:
Proposition 4.6. Let X C ]Rm+l be a connected, compact, boundaryless,
differentiable manifold of dimension m, and let f : X -+ §m be a continuous
mapping. Then X is the boundary of a bounded open set D c ]Rm+l, and

deg(j) = w(j, 0).

Proof. The Jordan Separation Theorem for compact hypersurfaces (III.6.2,


p. 124, and III.6.4, p. 129) says that X bounds a bounded open set D
as claimed. On the other hand, since both deg and ware invariant by
homotopy, we can suppose f is smooth, and then, by the smooth Tietze
Extension Theorem, f has a smooth extension J : D -+ ]Rm+l. Finally,
we pick a regular point a of J such that Iiall < 1. By IV.4.2, p.156,
w(j,O) = w(j, a); hence we are reduced to showing that
deg(j) = w(j, a).
4. Winding number 159

But in this situation w(f, a) = L:xE/-1(a) signxU), and from IVA.5, p. 157,
we get w(f, a) = deg(g), where

f(x) - a
9 : X --+ §m : x H ~.:.......:------::-
Ilf(x) - all·
Finally, the homotopy
f(x) - ta
Ht(x) = Ilf(x) - tall
is well defined, and by 111.304, p. 112, we conclude that

deg(f) = deg(Ho) = deg(H1 ) = deg(g) = w(f, a). o

Exercises and problems

Number 1. ([Poincare 1886] and [Bohl 1904]; 1.2, p. 19) Let X be a compact smooth
hypersurface of jRm+1 and let

f = (h, ... ,jm+1), 9 = (gl, ... ,gm+1) : X -+ jRm+1

be two continuous mappings with 0 tJ. f(X),g(X). Prove:


(1) If w(j, 0) '" w(g, 0), then there is at least one point x E X such that

hex) _ ... _ fm+1(x) < 0


gl(X) - - gm+1(X) .

(2) If w(j, 0) '" (_1)m+1w(g, 0), then there is at least one point Y E X such that

hey) = ... = fm+1(Y) > o.


gl(Y) gm+1(Y)

(3) If w(j,O) '" ±w(g,O), then the function hg1 + ... + fm+1gm+1 cannot have
constant sign on X.

Number 2. ([Kronecker 1869a] and [Hadamard 1910]; I.1, p. 10, and 1.2, p. 19) Let
f : X -+ jRm+1 be a smooth function on a compact smooth hypersurface X of jRm+1 ,
such that 0 tJ. f(X). Prove the following assertions:
(1) w(j,O) = deg (m).
(2) Denote by 0 the volume element of §m, and set rex) = x/llxll. Then

w(j,O) = vOl(~m) Ix (r 0 frO.

(3) Let x = (Xl, ... ,xm ) be local coordinates on X. Then

1 { -m-1 ( 8f(x) 8 f (x))


w(j,O) = vol(§m) Jx IIf(x)1I det f(x), 8X1 , ... , 8X m dx.
160 IV. Degree theory in Euclidean spaces

Number 3. ([Cauchy 1855]; 1.1, p.6) Fix a E C == ]R2. Let f : §1 -t C \ {a} be a


smooth mapping that has a holomorphic extension to some open neighborhood of §1 in
C. Prove the Argument Principle:

w(f, a)
1
= 27ry-1
r-1
1§1
f'(z)
f()
z - a
dz.

Check in this case the geometrical idea that the winding number tells how many times f
wraps §1 around the point a (1.1, p. 5): Given a half-line from the origin that meets f(§l)
transversally at finitely many points, w(f,O) is the difference between the number of them
at which the argument of f increases and the number of them at which the argument of
f decreases.

Number 4. ([Gauss 1813]; 1.1, p. 11) Let S C ]R3 be a compact surface. Compute
the winding number at a point a If. S of the inclusion S C ]R3, and deduce the Gauss
Theorem: Let v(x) = 11:~~13 be the electric field generated by a unit charge placed at a
point a. Then the flow of v through S is

1( ).-. -
s
V,Vs "s-
{47r
0
if a is inside S,
if a is outside S

(vs and Os stand for a normal vector field and the corresponding volume element of S).

Number 5. [Gauss 1833] Let f,g : [0,1] -t ]R3 be two smooth disjoint knots (f(0) =
f(l), g(O) = g(l), f(s) i- g(t) for all s, t). Consider the torus T C ]R3 parametrized by

T: x = (2 + cos(27rt)) cos(27rs), Y = (2 + cos(27rt)) sin(27rs), z = sin(27rt),

and define h : T -t ]R3 : (x, y, z) f-t f(s) - g(t). Prove that

w(h,O) = - 4~ 10 1 10 1 IIf(s) - g(t) 11- 3 det(f(s) - g(t), J' (s), g' (t))dsdt,

which is the link number of the two knots (1.2, p. 28).


Number 6. Let A be an orthogonal matrix of order m + 1. Use a winding number for
the computation of the degree of the smooth mapping f : §m -t §m : x f-t Ax. Also do
the direct computation in the Brouwer-Kronecker definition. Which is quicker?

5. The Borsuk-Ulam Theorem


Our primary goal here is to obtain the important Borsuk-Ulam Theorem.
Recall that a mapping f defined in some subset of a Euclidean space is
called even if f(x) = f( -x) and odd if f(x) = - f( -x). The famous
Borsuk-Ulam Theorem concerns the degree of such mappings. Afterwards,
we will deduce several purely topological results, including the Invariance
of Domain Theorem.
We start with a technical statement:
5. The Borsuk- Ulam Theorem 161

Lemma 5.1. (1) Let K c M be compact subsets of~n and let f : K ~ ~m,
m > n, be a continuous function such that 0 ct. f(K). Then f has a
continuous extension J: M ~ ~m such that 0 ct. J(M).
(2) Let D c ~n be a bounded open set, and let X = D \ D. We suppose
that 0 ct. D and D is symmetric with respect to 0 (that is, x f---t -x induces
a homeomorphism on D). Now let f : X ~ ~m, m > n, be an even (resp.,
odd) continuous mapping such that 0 ct. f(X). Then f has an even (resp.,
odd) continuous extension J: D ~ ~m such that 0 ct. J(D).
(3) Assume the same hypotheses as in (2), except that here m = n.
Then f has an even (resp., odd) continuous extension J : D ~ ~n such
that 0 ct. J(D n {x n = O}).

Proof. (1) Henceforth, norms of mappings are always intended on the com-
pact set M. Let e = dist(O, f(K)). First, by the Tietze Extension Theorem,
f extends to a continuous mapping 9 : ~n ~ ~m, and by the Weierstrass
Approximation Theorem there is a polynomial mapping h = h(x) such that
IIg - hll < le. Since m > n, the Easy Sard Theorem (II.3.4, p. 63) says
that the image h(~n) has empty interior in ~m, and we can find a point
a ct. h(~n) such that Iiall < le.
Now consider h' = h - a : ~n ~ ~m.
Clearly 0 ct. h' (~n), and furthermore

II h' - gil::; II h' - h II + II h - gil < ! e.


Next consider the continuous mapping
Eh'(x)
h"(x) = { 2I1h'(x) II
for IIh'(x)1I ::; !E,
h'(x) otherwise.

We have IIh"lI ~ !E. On the other hand, f(x) = g(x) for x E K, and
IIh'(x)1I ~ IIg(x)II-lIg(x) - h'(x) II > E- !E = !E,
so that h"(x) = h'(x), and we conclude

IIh"(x) - f(x) II <!E for x E K.


Now, we can apply the Tietze Extension Theorem to the function
h" - f: K ~ B,

where B is the open ball of radius !E


(homeomorphic to ~m), and h" - f
extends to a continuous mapping h'" : M ~ B. We claim that J = h" - h'"
is the mapping we seek. Indeed, for x E K we have

J(x) = h"(x) - h"'(x) = h"(x) - (h"(x) - f(x)) = f(x),


162 IV. Degree theory in Euclidean spaces

and taking norms on M,

Ilfll ~ IIh"II - IIh"'II ~ !E - IIh"'II > !E - !E = a.


This completes the proof of (1).
The proof of (2) goes by induction on n. For n = 1 the set D is a union
of disjoint open intervals 1 = (a, b) with a > a and their symmetric copies
-1 = (-b, -a), and X consists of the end points of those intervals. To
extend f to D, we map each 1 onto a polygonal path in jRm \ {a} joining
f(a) to f(b) (here we use m > 1), and for x E -1 define f(x) = f( -x) in
the even case and f(x) = -f(-x) in the odd case.
Now suppose the result holds for n-l. We identify jRn-1 == {x n = a} c
jRn and consider the set
D1 = D n jRn-1,
which is symmetric. Clearly, the set Xl = D1 \ D1 C jRn-1 is contained
in X, and by induction we find a continuous mapping f1 : D1 -+ jRm that
extends h = fix! and is never a. Now, since D1 n X = 0, we can define a
nowhere zero continuous mapping 12 : D1 U X -+ jRm by

{ 12 ~ f1 on D1,
12 = f on X.

After this preparation we can apply (1). We already have one compact set,
namely
K=D1 UX,
and to define M :J K, set

D+ = D n {x n > a}, D_ = D n {x n < a},

and take
M = D1 U X U D+ = D1 U xu D+.
Now (1) gives a continuous mapping f2 : M -+ jRm that extends 12IK and is
never a. To conclude the proof of (2), one extends f2 to D_ by symmetry.
Finally, the proof of (3) is analogous, but one uses (2) instead of in-
duction. The argument ends with an application of the Tietze Extension
Theorem instead of (1), which gives the condition a rJ. f(D n jRn-1) instead
of a rJ. f(D). 0

After these extension lemmas, we are ready to prove the famous Borsuk-
Ulam Theorem:
5. The Borsuk-Ulam Theorem 163

Theorem 5.2 (Borsuk-Ulam Theorem). Let D c IRn be a symmetric


bounded open set, with 0 E D. Set X = D \ D, and let f : X ---+ IRn
be a continuous mapping with 0 ¢ f(X).

(1) Assume
f(x) f(-x)
Ilf(x)11 =1= Ilf( -x)11 for all x E X.
Then the winding number w(f, 0) is odd, hence not zero.

(2) Assume
f(x) f( -x)
Ilf(x)11 =1= -llf( -x)11 for all x E X.
Then the winding number w(f, 0) is even, and zero for n odd.

In particular, (1) applies if f is an odd function, and (2) applies if f is an


even function

Proof. We start with case (1). By the Tietze Extension Theorem, f extends
to a continuous function on D, again denoted f : D ---+ IRn. But we need
that extension to be odd. To amend that, consider the function

9 : D ---+ IRn : X f---t f(x) - f( -x),

which is clearly an odd function. Then we define the homotopy

Ht(x) = f(x) - tf( -x), Ho = f, HI = g.


We claim that 0 ¢ H([O, 1] x X). Indeed, suppose 0 = f(x) - tf( -x) for
some x E X, 0 ~ t ~ 1. Since 0 ¢ f(X), it must be true that t > 0,
and from f(x) = tf( -x) it follows that II~~:~II = II~~=:~II' contrary to the
hypothesis.
We conclude that f and 9 have the same degree, and consequently we
can substitute 9 for f, or merely suppose that f : D ---+ IRn is odd, which
we do henceforth.
Pick c > 0 small enough so that U = {llxll ~ c} c D and define the
mapping
f 1: UUX TT'Iln
---+~ :Xf---t
{fx
(x) if x EX,
ifxEU.

Set DI =D\U, Xl =DI \D}, and /2=hlxl; note that Xl =xu{llxll =c}.
The following figure depicts the setting:
164 IV. Degree theory in Euclidean spaces

~n-l

We can apply IV.5.1, p. 161, and obtain an odd continuous extension


12 : Dl --+ ~n of 12, such that 0 ~ 12(D 1 n ~n-l). Next, we set

f 3: D --+IN..
llJln {12(X) if x E D 1 ,
:xt---+
x if x E U.

It is clear that hlx = fix; hence by IV.2.6, p. 150,

d(f, D, 0) = d(h, D, 0).

Now we will use the additive property of the degree, IV.2.5(3), p. 149.
We write (see the figure above)

D1+ = Dl n {x n > O}, D 1- = Dl n {xn < O}.

As 0 ~ 12(D 1 \ D1+ U D 1 -), from additivity it follows that

We make the change of variables (a linear isometry in fact) <p : D1+ --+
Dl- : x t---+ -x, to get

Then, by direct computation,

and consequently,
5. The Borsuk-Ularn Theorem 165

To conclude, we look at the interior Uo = {llxll < c-} of U. We have


D => Uo U (D \ U), and 0 ff- h(D \ Uo U (D \ U)), which again by additivity,
gives

d(f, D, 0) = d(h, D, 0) = d(hl uo ' Uo, 0) + d(f3, D \ U,O)


= d(f3, Uo, 0) + d(f3, D I , 0)
= d(Id, Uo, 0) + d(J2, D I , 0) = 1 + 2N,
which is an odd number, as desired.
Thus we have proved part (1) of the theorem. For (2), the same argu-
ment works, with the modifications we describe next.
To reduce to the case where f is defined on the whole D, the function
9 needed is g(x) = f(x)+ f(-x). The function II must be defined as

f 1: U U X -tJN..
1[])n {f(X) if x E X,
:XH
Ixl = (IXII, .. ·, Ixnl) if X E U.

Accordingly,

Then,

Thus,

and consequently,

Finally

d(f, D, 0) = d(f3, D, 0) = d(h, Uo, 0) + d(f3, D \ U,O)


= d(f3, Uo, 0) + d(f3, D I , 0)
= d(1 Id I, UO, 0) + d(J2, DI, 0) = 0 + (1 + (-1)n)N,
which is the required number in part (2). Note here that d(1 Id I, UO, 0) = 0
by IV.2.5(2), p. 149. Indeed, d(1 Id I, UO, 0) = d(1 Id I, UO, a) = 0 for a =
(-c-,O, ... ,O) ff-IIdl(X). 0

The above theorem has a very nice converse:


166 IV. Degree theory in Euclidean spaces

Theorem 5.3 (Hirsch Theorem). Let D c ~n be a symmetric bounded


open set, with 0 ED. Set X = D \ D, and let f : X ~ ~n be a continuous
mapping with 0 1. f (X) .

(1) If the winding number w(j, 0) is odd, then there is some x E X such
that
f(x) - f( -x)
IIf(x)11 IIf( -x)II'
(2) If the winding number w(j, 0) is even, then there is some x E X such
that
f(x) f(-x)
Ilf(x)1I Ilf( -x)II'
Proof. (1) Assume by way of contradiction that

f(x) -f(-x)
Ilf(x)1I of: Ilf(-x)11
for all x EX. We extend f to a continuous mapping 1 : D ~ ~n and
define a homotopy by

Ht(x) = (1 - t)J( -x) + tJ(x).


By assumption, Ht(x) of: 0 for x E X, and by IV.2.4, p. 147,

w(j,O) = d(/,D,O) = d(Hl,D,O).


2

Now note that


Hdx) = !J(-x)
2
+ !J(x)
is an even mapping, and by the Borsuk-Ulam Theorem, IV.5.2(2), p. 163,
the degree d(H 1, D, 0) must be even. This is a contradiction, because
2
w(j,O) is odd.
Part (2) is proven similarly: one assumes

f(x) f( -x)
Ilf(x)1I of: IIf(-x)11
for all x EX, extends f to 1, and defines
Ht(x) = (1 - t)J( -x) - tJ(x).

Then the odd mapping H 1 has even degree, contrary to the Borsuk-Ulam
2
Theorem, IV.5.2(1), p. 163. 0
5. The Borsuk- Ulam Theorem 167

Note that from the Borsuk-Ulam Theorem we see readily that a function
like f in case (1) on p. 163 (for instance an odd J) must have some zero in
D after any extension. It is also easy to obtain a fixed point result:
Corollary 5.4. Let Dc ]Rn be a symmetric bounded open set, with 0 E D.
Set X = D \ D, and let f : D -+ ]Rn be a continuous mapping whose
restriction fix is odd. Then there is some x E D such that f(x) = x.

Proof. The function 9 : D -+ ]Rn : x f---+ x - f(x) also has an odd restriction
glx:
g( -x) = -x - f( -x) = -x + f(x) = -g(x) for x E X.
If 0 E g(X), we are done, so we assume 0 tJ. g(X). Then, by the Borsuk-
Ulam Theorem, IV.5.2(1), p. 163, d(g, D, 0) =I 0; hence there is some xED
such that g(x) = 0 and f(x) = x. 0

Here is a second interesting consequence:


Corollary 5.5. Let Dc ]Rn be a symmetric bounded open set, with 0 E D.
Set X = D \ D, and let f : X -+]Rm be a continuous mapping with m < n.
Then there is an x E X such that f(x) = f( -x).

Proof. Set ]Rm == {Xm+l = ... = Xn = O} C ]Rn, and consider the odd
mapping 9 : X -+ ]Rm : x f---+ f(x) - f( -x). We assume 0 tJ. g(X) (otherwise
there would be nothing to prove). By the Tietze Extension Theorem 9
has a continuous extension 9 : D -+ ]Rm, and we then define 9 : D -+
]Rn : x f---+ (g(x),O). Like g, the preceding function is odd on X, and by
assumption 0 tJ. g(X). Thus, by the Borsuk-Ulam Theorem, IV.5.2(1), p.
163, the degree of 9 is not zero: d(g, D, 0) =I O. However, pick c > 0 small
enough so that af; = (0, ... ,0, c) E ]Rn \ g(X). Then, by IV.2.3, p. 146,

d(g, D, 0) = d(g, D, af;),


and the degree in the right-hand side is zero, because af; tJ. g(D) (IV.2.5(2),
p. 149).

This contradiction comes from our assumption that 0 tJ. g(X); thus
there must be some x E X with f(x) = f( -x). 0

Now we give a very important theorem (which for dimension 1 is an


immediate consequence of the Bolzano Theorem):
Theorem 5.6 (Invariance of Domain Theorem). Let W c ]Rn be an open
set, and let 9 : W -+]Rn be a locally injective continuous mapping. Then 9
is an open mapping.
168 IV. Degree theory in Euclidean spaces

Proof. Let a E Wand consider the translated open set

D =W - a = {x - a: x E W}

(which contains 0) and the continuous mapping

f :D ~ ~n : x ~ g(a + x) - g(a).

By hypothesis, for c > 0 small enough, f is injective on the closure B of


the open ball B = {llxll < c} c Dj we set X = B \ B. Then we define the
following homotopy:

Ht(x) = fe: t) - f(;~Xt)' x E B.

We see that 0 rt H([O,lJ x X): if H(t, x) = 0, since fiB is injective,


l~t = ~!~ and x = 0 rt X. Thus, by homotopy invariance, IV.2.4, p. 147,

d(f, B, 0) = d(Ho, B, 0) = d(Hl' B, 0).


But the mapping
Hl(X) = f(!x) - f(-!x)
is odd, and by the Borsuk-Ulam Theorem, IV.5.2(1), p. 163, its degree is
not zero. Let V denote the connected component of~n\f(X) that contains
O. We deduce (IV.2.3, p. 146) that

d(f, B, y) =1= 0 for every y E V,

which by IV.2.5(2), p. 149, means y E f(B). Consequently,

o EVe f(B).
Thus, there is an "1 > 0 such that
if Ily'll < "1, then y' = f(x') for some x' E ~n with IIx'll < c.
Setting x' = x - a and y' = y - g(a), we deduce that
if Ily - g(a) II < "1, then y = g(x) for some x Ilx - all < c.
E ~n with
This means that g(a + B) is a neighborhood of g(a), and since the a + B's
for c small enough form a neighborhood basis of a, we conclude that 9
maps every neighborhood of a onto a neighborhood of g(a). Since a E W
is arbitrary, 9 is open. D
5. The Borsuk-Ulam Theorem 169

Remark 5.7. Suppose W = ~n in the above theorem. Then, if

lim IIf(x)11 = +00,


lixll-Hoc
the mapping f is surjective.
Indeed, that condition implies that f is proper, and then f(~n) is not
only open as the theorem says, but also closed in ~n. D

Variations of the previous theorem follow:

Corollary 5.8. (1) (Invariance of Domain Theorem) Let W c ~n be an


open set, and let g : W --+ ~n be an injective continuous mapping. Then
V = g(W) C ~n is open, and g is a homeomorphism onto V.
(2) (Invariance of interiors) Let S, T c ~n be arbitrary sets, and let
g : S --+ T be a homeomorphim. Then

(of course, Intn means interior in ~n).


(3) (Invariance of dimension) Let W c ~n and V c ~m be non-empty
open sets. If Wand V are homeomorphic, then n = m.

Proof. Assertion (1) is an immediate consequence of the previous theorem.


For (2), note that since f is injective, f(Intn(S)) is an open subset of~n;
hence f(Intn(S)) C Intn(T). Arguing for f- 1 , we get the other inclusion.
Equality concerning closures follows by complementation, because

A n ~n \ A = A \ Intn(A) for any A.

For (3) suppose n < m, and put W' = W x {O} C ~n X ~m-n = ~m.
Of course, W' is homeomorphic to W, hence to V by hypothesis. Then, by
(2), Intm(W') is homeomorphic to Intm(V) = V. But this is impossible,
because W' C {xm = O} has empty interior in ~m. D

We conclude the section with a common improvement to the different


fixed point theorems seen so far (III.7.2, p. 134, and IV.2.9, p. 152):

Proposition 5.9. Let D C ~n be a bounded open set whose closure D is


homeomorphic to a closed ball, and denote X = D \ D as usual. Then every
continuous mapping f : D --+ ~n such that f(X) C D has some fixed point.
170 IV. Degree theory in Euclidean spaces

Proof. Let h : D -+ yr be a homeomorphism onto the unit closed ball yr c


]Rn; by IV.5.8(2), p. 169, h(X) is the unit sphere §n-l. Since h(f(X)) c
h(D) = yr, we can define a homotopy Ht : X -+]Rn by

Ht(x) = x - h-1(th(f(x))), 0 ~ t ~ 1.

Now suppose Ht(x) = 0 for some x E X. Then x = h-l(th(f(x))) E X, so


that
th(f(x)) = h(x) E §n-l
and as h(f (x)) E yr, it must be true that t = 1. Consequently,

x = h-l(th(f(x))) = h-l(h(f(x))) = f(x),

and we have a fixed point in X. Otherwise, we in fact have H t : X -+


]Rn \ {O} and by the definition and homotopy invariance of winding numbers,

d(Id - f, D, 0) = w(Id - fix, 0) = W(Hl' 0) = w(Ho, 0)


= w(Id-h-1(0)lx,0) = d(Id-h-1(0),D,0) = 1,

because h-1(0) E D. Thus, d(Id - f, D, 0) i= 0, and by IV.2.5(2), p. 149,


there is an xED such that x - f(x) = 0; that is, we have a fixed point in
D. D

Exercises and problems

Number 1. Suppose we have a decomposition

into non-empty closed subsets with Ai n -Ai = 0.


(1) Show that for every i = 1, ... , m + 1 there is an odd continuous function Ii :
§"' --+ ]R such that J;(x) = 2 if and only if x E Ai and J;(x) = -2 if and only if x E -Ai.
(2) Deduce that n::i 1 Ai t= 0.

Number 2. Let AI, ... , A m+ 1 be non-empty closed subsets ofthe sphere §m. Prove the
following:
(1) If §m = Al u··· U A m+ l , there is a point x E §"' such that x E Ai and -x E Ai
for some i.
(2) If §m = (AI U ... U Am) U (-AI U··· U -Am), there is a point x E §"' such that
x E Ai and -x E Ai for some i.
Number 3. Let /1, ... , 1m : ]Rm+l --+ ]R be homogeneous continuous mappings of odd
degree (that is, li()..X) = )..Pi Ii (X) with Pi odd). Prove that the system /1 = 0, ... ,1m = 0
has some non-trivial solution.
Number 4. Let I : ]Rn --+ ]RP be a homogeneous continuous mapping of odd degree,
without zeros. Prove that every linear subspace L c ]RP of dimension n - 1 is perpendic-
ular to I(x) for some x E ]Rn.
6. The Multiplication Formula 171

Number 5. An odd continuous mapping f : sm -+ sn is also called equivariant, ex-


pressing that it commutes with the antipodal involution. Prove the following:
(1) There are equivariant mappings f : sm -+ sn if and only if m S n.
(2) Every equivariant mapping f : sm -+ sm has odd degree; hence it is not null-
homotopic.
(3) Every equivariant mapping f : sm -+ sn, m < n, is not surjective, hence null-
homotopic.
Number 6. Here we consider topological spaces X equipped with a continuous invo-
lution ax (that is, ax 0 ax = Idx); in particular, asm is the antipodal involution. A
continuous mapping f : X -+ Y of spaces with involution is called equivariant when
ay 0 f = f 0 a x. For a space with involution X, define the following:
(i) the level s(X), as the infimum (possibly 00) of all integers m ~ 1 such that there
is an equivariant continuous mapping X -+ sm,
(ii) the colevel s'(X), as the supremum (possibly 00) of all integers m ~ 1 such that
there is an equivariant continuous mapping sm -+ X.

Prove that s'(X) S s(X). Compute both invariants when X is the boundary of
a symmetric bounded open neighborhood D of the origin in Rn, equipped with the
antipodal involution.

Number 7. Consider the polynomial algebra A = R[Xl, ... ,xn ]/(l + x~ + ... + x~).
Complete the details of the following proof that -1 is not a sum of less than n squares
in A.
(1) Otherwise, let -1 = h (X)2 + ... + f~-l (x) + fo(x)(l + x~ + ... + x~) for some
polynomials A(x) E R[x], and write

fk( yCI x) = Pk(X) + yCIqk(X), Pk(X), qk(X) E R[x], Pk(X) even and qk(X) odd.
(2) Replacing x with A x in the equation for -1 above and comparing real parts,
obtain
n-l
-1 = I)Pk(x)2 - qk(X)2) + po(x)(l - xi - ... - x~).
k=l

(3) Show that the mapping q = (ql, ... , qn-d : D n -+ R n - 1 collapses two antipodal
points in aD n = sn-l, and being odd, it has a zero a E sn-l.
(4) Substituting x = a in the equation for -1 in (2), get a contradiction.
Number 8. Invariance of Domain holds for manifolds, but no further:
(1) Prove that a locally injective continuous mapping f : M -+ N of boundaryless
manifolds of the same dimension is open.
(2) Consider the lemniscate X C R2 with polynomial equation (X2 + y2)2 = xy.
Define a continuous bijection f : R -+ X which is not a homeomorphism.

6. The Multiplication Formula


In this section we prove the general formula for the computation of the
degree of a composite mapping. This is much more delicate than the corre-
sponding resuit for the Brouwer-Kronecker degree, but always elementary.
Here is the statement of the formula:
172 IV. Degree theory in Euclidean spaces

Proposition 6.1 (Multiplication Formula). Let D, E c jRn be bounded


open sets with boundaries X = D \ D and Y = E \ E. Let f : D -+ jRn and
9 : E -+ jRn be continuous mappings such that f(D) C E. The open set
E\f(X) decomposes into countably many connected components El, and we
arbitrarily choose one point Cl E El in each. Now, let a E jRn\g(YUf(X)).
Then
d(g 0 f, D, a) = L
d(g, El, a)· d(f, D, Cl).
l

Proof. We split the argument into several steps.


Step I: The formula is well defined. First of all, the degree d(g 0 f, D, a)
exists, because jRn \ g(f(X)) ~ jRn \ g(Y U f(X)). Next, since El is a
connected subset of E\f(X) C jRn\f(X), the degree d(f,D,·) is constant
on it (IV.2.3, p. 146). For the other factor d(g, El, a) we must see that
a rJ- g(El \ El), but we have

El \ El c (E \ E) U f(X) = Y U f(X)

and a rJ- g(YUf(X)). Finally, the sum Ll is finite. Indeed, since g-l(a) C
E \ f(X) is a compact set, it is contained in the union of finitely many
Ee's, which are disjoint; hence g-l(a) n El = 0 for, say, f ~ fo. Thus, by
IV.2.5(2), p. 149, d(g, Ee, a) = 0 for f ~ fo.
Step II: Approximation data. By the Weierstrass Approximation Theorem,
there is a polynomial mapping j such that on D

Ilf _ jll < {dist(f(D),jRn \ E) and


dist(g-l(a), f(X)).

From the first bound we deduce that j(D) C E, and the composite map-
ping go j does exist. From the second bound, we see that a rJ- g(j(X)),
hence a rJ-jRn \ g(Y U j(X)), and c = dist(a,g(Y U j(X))) > O. Again by
the Weierstrass Approximation Theorem, there is a polynomial mapping
gl such that IIg - gIll < !c on E, and by the Sard-Brown Theorem, there
is a regular value b of the mapping gl 0 jlD with lIa - bll < !c. We consider
now the polynomial mapping g = gl + a-b. We have:

(i) On E,
IIg - gil ~ IIg - gIll + IIgl - gil < !c + !c = c;
hence a E jRn \ g(Y U j(X)).

(ii) By translation, a is a regular value of g 0 jlD.


6. The Multiplication Formula 173

Step III: Reduction to the approximation data. We claim that

d(g 0 f, D, a) = d(g 0 j, D, a) = d(9 0 j, D, a).

For the first equality, we consider the homotopy

4>t(x) = (1- t)f(x) + tj(x), x E D.

We have

Ilf(x) - 4>t(x) I = tllf(x) - j(x) II < dist(f(D),lRn \ E);

hence 4>t(x) E E, and the homotopy go 4>t is well defined. Furthermore,


art g(4)([O, 1] x X)). Indeed, suppose there is an x E X such that 4>t(x) E
g-l(a). Then

l14>t(x) - f(x) II = tllj(x) - f(x) I < dist(g-l(a), f(X)),

which is impossible. Thus, we can use homotopy invariance (IV.2.4, p. 147)


and

d(g 0 f, D, a) = d(g 0 4>0, D, a) = d(g 0 4>1, D, a) = d(g 0 j, D, a).

For the second equality of our claim we use the homotopy

llit(x) = (1 - t)g(j(x)) + tg(j(x)), x E D.

Again, the key fact is that a rt lli([O, 1] x X). But if x E X, we have

lIa - llit(x) I = I (a - gj(x))


+ t(gj(x) - gj(x)) II
2: lIa - gj(x)lI- tllgj(x) - gj(x) I
2: dist(a, g(j(X))) -lig - gil > dist(a, g(j(X))) - c 2: 0,
by (i) in the preceding step. Thus, we can indeed apply homotopy invari-
ance.
Step IV: Multiplication. By IV.2.1, p. 145, we know that

d(g 0 f, D, a) = L signx(g j) = L signj(x) (g) signx(j)


0

xE(fjo/)-l(a) xE(fjoj)-l(a)

= L L signy(g)· signx(j)
YEfj-l(a) xEj-l(y)
= L signy(g) ( L
Signx(j)) = L signy(9). d(j, D, y).
yEfj-l(a) xEj-l(y) yEfj-l(a)
174 IV. Degree theory in Euclidean spaces

Now, for every k the set

Wk~ = {y E E \ j(X): d(j,D,y) = k}


is a union of connected components of E \ j(X), whose boundaries are
contained in Y U j(X) (as for the Ep's). Hence, by (ii) in Step II, a rt
9(Wk~\ Wk), and we can continue the above sequence of equalities as follows:
d(goj,D,a) = ...
= L signy(§)' d(j, D, y)

= Lk( L signy(g)) = Lk. d(g, Wk~a).


k yEg-l(a) nWk- k

Step V: Second multiplication. We repeat the same computation for j. We


set
Wk = {y E E\j(X): d(j,D,y) = k} = UEf;,
i

where the Efi's are the connected components of E\j (X) on which d(j, D, .)
== k. Since W k \ Wk c YUj(X), we have art g(W k \ Wk), and by additivity
(IV.2.5(3), p. 149) we get

d(g, Wk,a) = Ld(g,Efi,a).


i

Indeed, as remarked in Step I, the sum is in fact finite:

L d(g, Efi' a) = L d(g, Efi' a)


i i: fi<eO

because g-l(a) n Ef = 0 for f 2: fa; moreover, in this situation

art 9(Wk \ Wk) = 9(Wk \ U Eei)


i: ei<fo

and the additivity property does apply.


Notice now that the Eei's in this sum are the Ee's contained in Wk. We
conclude that

L d(g, Ee, a) . d(j, D, ce) = L k· L d(g, Ee, a)


f k f:EecWk

= Lk. d(g, Wk,a).


k
6. The Multiplication Formula 175

Step VI: Simplification of mappings. Comparing the formulas in the pre-


ceding two steps, we are reduced to showing that

for every k. But by (i) in Step II,

Ilg - gil < E = dist (a,g(YU j(X))) :S dist (a,g(W;\ W k-)).


Hence

and we must show the equality of the following two degrees:

d(g, Wk, a) = d(g, Wk~ a).

Step VII: Computation of the left-hand side degree. We claim that

1- g(Wk \ W;).
a

Suppose a = g(z) with z E Wk \ Wk~ Since a 1- g(YUf(X)) and Wk \ Wk C


Y U f(X), it must be true that z E Wk, so that dU, D, z) = k. Moreover,

II f - jll < dist(g -1 (a), f(X)) :S dist(z, f(X))


and we get d(j, D, z) = dU, D, z) = k, that is, z E W;, a contradiction.
Thus the claim is proved, and, by additivity (IV.2.5(3), p. 149),

d(g, Wk, a) = d(g, W; n Wk, a).

Step VIII: Computation of the right-hand side degree and conclusion. The
argument is analogous. First, a 1- 9(Wk-\ W k ), because a 1- g(Y U j(X))
and Wk-\ Wk- C Y U j(X). Then, additivity gives

d(g, Wk-,a) = d(g, Wk-n Wk,a).

Thus, we have the equality of degrees stated at the end of Step VI, and
the proof of the Multiplication Formula is finished. 0

Exercises and problems

Number 1. Find two continuous functions f : D --+ ]Rn and 9 : E --+ ]Rn verifying the
hypothesis of the Multiplication Formula (IV.6.1, p. 172), with D and E connected, such
that the sum in that formula has at least two different summands.
176 IV. Degree theory in Euclidean spaces

Number 2. Let D (resp., E) be a bounded open subset ofli", and let I: I5 -+ lin (resp.,
9 : E -+ lin) be a continuous function. Suppose E is connected and that I(D) c E,
I(8D) C 8E. Prove that
d(g 0 I, D, a) = d(g, E, a) . d(j, D, b)
for a E lin \ g(8E) and bE I(D).
Number 3. Let D (resp., E) be a bounded open subset of lin , and let I: D -+ lin (resp.,
9 : E -+ lin) be a continuous function. Suppose D is connected and that I(8D)nI(D) = 0
and I(D) C E. Prove that

d(g 0 I, D, a) = d(g, E, a) . d(j, D, b)


for a E lin \g(8EU I(8D)) such that g-l (a) C I(D) and bEE. Show that the condition
g-l(a) C I(D) is necessary for this conclusion.

°
Number 4. For p> we denote by D~ C lik the open ball of radius p centered at the
origin and byD: and S;-l the corresponding closed ball and sphere; if p = 1, we just get
Dk, Dk, and §k-l. Fix n 2: 1, and set Ep = (-p,p) x D;.
(1) Compute the degree of the mapping

!pp: Ep -+ n;+l : (t,x) t--+ (t, ~Jp2 - t 2 x).


(2) Show that for every continuous mapping I: D n -+ lin such that J(§n-l) C §n-l
and (Id[-l,l) Xf)(El) C Er with r > 1,

d(j,D n , 0) = d(!Pr 0 (Id[-l,l) xf),El,O).


(3) Prove that for every continuous mapping 9 : §n-l -+ S',-l there is a unique
continuous mapping E(g) : §n -+ §n (the suspension of g) such that

!PI 0 (Id[-l,l) xg) = E(g) 0 !PI,


and prove that deg(E(g)) = deg(g).
Use the above construction to produce by induction on k a mapping §k -+ §k of any
given degree d.

7. The Jordan Separation Theorem


The core of this section is to prove the following purely topological state-
ment:

Theorem 7.1 (Jordan Separation Theorem). Let K, L c ]Rn, n ~ 2, be


two compact sets. If K and L are homeomorphic, then their complements
]Rn \ K and]Rn \ L have the same number N ::; 00 of connected components.

Proof. Note first that both complements ]Rn \ K and ]Rn \ L are open sets
with exactly one unbounded connected component each, which we denote
7. The Jordan Separation Theorem 177

£1 0 and To, respectively. In fact, we consider the decompositions into con-


nected components

IRn \ K = Ui2:0 £1i' a = number of bounded £1i's :::; 00,


{
lRn \ L = U j 2:0 Tj , (3 = number of bounded Tj's :::; 00.

Pick points ai E £1 i (i ~ 1) and bj E Tj (j ~ 1) . Also, we consider as usual


the boundaries

By hypothesis, we have a homeomorphism h : K ---+ L, and by the Tietze


Extension Theorem we find continuous mappings cp, 'l/J : lRn ---+ lRn such that
cplK = hand 'l/JIL = h- 1 and both cp and 'l/J are the identity off an open ball
containing K and L. In this situation, we easily compute the degrees

Indeed, by construction

and

Finally, since cp 0 'l/JIYi = Id h'i and 'l/J 0 cplxi = Id lXi' by the Boundary
Theorem, IV.2.6, p. 150,

d( cP 0 .1. 1"'
'f" 1. i,
b)
j = d(Id ,r,i , bj ) = {I if bj E ri, that is, if i
0 otherwise
=j } = 8.oJ'

and

Aaj) = d(Id , ..:...li,


d( 'l/J 0 cp, ..:...li, A aj ) = {I if aj E Lli' that is, if i =
o otherwise
j} = >:. '.
uoJ

Next, we compute d(cp 0 'l/J, Ti, bj) by using the Multiplication Formula.
Step I: Multiplication setting. We decompose the complement of the com-
pact set 'l/J(Yi) into connected components, the first one unbounded:

Now, this complement contains


178 IV. Degree theory in Euclidean spaces

hence each connected component of the last union is contained in a unique


De, and we can write
and Do \ Uo C K,
and Dl \ Ul C K,

Next we choose a big enough open ball E C ~n to have the following


conditions:

(i) bj tt. 'P(E \ E) (recall that 'P is the identity off an open ball).

(ii) E:J 'l/J(ri) U (~n \ Do); hence E :J Ue~l De .


(iii) Eo = En Do is connected, and Eo ct'l/J(ri).
(iv) E \ 'l/J(Yi) = Eo U El U E2 ... , where Ee = De for P 2: 1.

In this situation we can apply the Multiplication Formula, IV.B.1, p. 172, to


the mappings 'l/Jlri : r i -+]Rn and 'PIE: E -+]Rn to compute d('P °'l/J, ri, bj).
Step II: Computation by multiplication. Using the Multiplication Formula,
we obtain
d('P0'l/J,ri,bj ) = "Ld('P,Ee,bj)d('l/J,ri,Cf),
e~o

for any chosen Ce E Ee. Here we can immediately discard the P = 0 term
by picking CO E Eo \ 'l/J(ri) (condition (iii) above). Thus,

d('P ° 'l/J,ri,bj ) = "Ld('P,Ee,bj)d('l/J,ri,ce),


e~l

and Ce E De = Ee.
Step III: Additional expansion. Here we further decompose some terms of
the last sum. In fact, we are to show that

First, this is a well-defined finite sum, for, since

bj E rj C ~n \ L= ]Rn \ 'P(K) C ]Rn \ 'P(Xkr(f») ,


every degree d( 'P, Llkr(l) , bj ) exists. Furthermore it is zero for r big enough.
Indeed, since bj tt. L = 'P(K), 'P-1(bj ) n K = 0; hence

'P-l(bj) \ .10 c ULlk ·


k~l
7. The Jordan Separation Theorem 179

But 'P-l(bj) \ ..10 is compact (closed and bounded), so that for r large, say
r > ro, 'P- 1 (bj) n Llkr(f) = 0; hence d( 'P, Llkr(f) , bj) = 0, and
ro
Ld('P, Llkr(f) , bj) = Ld('P, Llkr(f),bj).
r r=l

Now, by construction we know that

bj ¢ 'P(Df \ ULlkr(e)),
r

and since 'P-l(bj ) does not meet Llkr(f) for any r > ro, we get
ro
bj ¢ 'P( Df \ ULlkr(e)).
r=l

Thus we can apply additivity (IV.2.5(3), p. 149), to see that the degree
d( 'P, Ee, bj ) is the finite sum above.
Step IV: End of the computation. Putting Steps II and III together, we
have

d('P 0 'ljJ, ri, bj ) =L ( L d('P, Llkr(f) , bj ) )d('ljJ, r i , Cf)


e~l r

=L d('P,Llkr(f),bj). d('ljJ,ri,akr(f))·
f,r~l

The last equality follows by rearranging the sums and taking Ce = akr(e) E
Llkr(f) C De for every e, i. But the last sum can also be rewritten as

... = L d('P,Llk,bj)·d('ljJ,ri,ak)
LlkrtDo

= L d('P, Llk' bj) . d('ljJ, r i , ak).


k~l

This is because if Llk c Do, then

for any ale E Do (IV.2.3, p. 146). But we can find ale far from 'ljJ(ri ) so that
the equation 'ljJ(x) = ale has no solution in r i and the last degree is zero
(IV.2.5(2), p. 149).
With this, we have finished the computation of d( 'P 0 'ljJ, ri, bj ).
180 IV. Degree theory in Euclidean spaces

All that has been done so far for r.po7jJ can be done analogously for 7jJor.p,
and one ends up with the following formulas:

(*) L d( r.p, Llk' bj) . d( 7jJ, n, ak) = d( r.p 0 7jJ, r i, bj ) = 6ij ,


k~l

(**) L d( 7jJ, re, aj) . d( r.p, Lli' be) = d( 7jJ 0 r.p, Lli' aj) = 6ij .
e~l

With these formulas, we can finally prove O! = (3.


To that end, we define two linear mappings. Let R(-r) stand for the
direct sum of 1 :S 00 copies of R, and let {ei} stand for its canonical basis.
Then

A: R(a) -+ RCB) : ek t---+ (d(7jJ, r i , ak) : i ~ 1) = (Aik : i ~ 1),


{
B : RCB) -+ R(a) : ee t---+ (d(r.p, Lli' be) : i ~ 1) = (Bie : i ~ 1).

Now, we can rewrite the formulas above as

""'
L.J Bk·J . A-k
t -- 6··
tJ' (**) L Aej . Bu = 6ij
k~l e~l

and we see that A and B are inverse to each other; hence both linear spaces
have the same dimension, and O! = (3, as desired. 0

Examples 7.2. (1) A compact set KeRn that is homeomorphic to a


differentiable hypersurface disconnects Rn.
(2) A compact set KeRn that is homeomorphic to a connected differ-
entiable hypersurface disconnects R n into two open connected components,
one bounded, whose boundaries are both K. No proper subset T c K
disconnects Rn.
(3) An arc (that is, a compact set homeomorphic to an interval) never
disconnects Rn for n ~ 2.
(4) A topological closed ball never disconnects R n for n ~ 2.
(5) A compact set KeRn homeomorphic to a proper subset of §n-l
does not disconnect Rn.
(6) A sphere is not homeomorphic to any of its proper subsets (oth-
erwise we would have two homeomorphic compact subsets of a Euclidean
space, one disconnecting it and the other not). 0
7. The Jordan Separation Theorem 181

The Jordan Separation Theorem, IV.7.1, p. 176, also holds when we


replace Euclidean spaces by spheres:
Proposition 7.3. Let K, L c §n, n ~ 2, be two compact sets. If K and
L are homeomorphic, then their complements §n \ K and §n \ L have the
same number N ~ 00 of connected components.

Proof. If one of the two compact sets is the sphere, then the other is the
sphere also, by IV.7.2(6) above. Hence we can pick two points a E §n \ K
and b E §n \ L and consider the stereographic projections from them, de-
noted by x : §n \ {a} -+ ~n and y : §n \ {b} -+ ~n, which are diffeomor-
phisms. Then x{K) and y{L) are compact homeomorphic subsets of ~n;
hence they disconnect it in the same number of connected components.
Pulling them back to §n via x and y, we see that §n \ K and §n \ L have the
same number of connected components too. Indeed, it suffices to remark
that the unbounded component U of ~n \ x{K) (resp., V of ~n \ y{L))
corresponds to the connected component U' = {a} U x-I (U) of §n \ K that
contains a (resp., V' = {b} U y-I{V) of §n \ L that contains b). 0

We can finally generalize IV.7.1, p. 176, to closed sets:

Proposition 7.4. Let G, G' c ~n, n ~ 2, be two closed sets. If G and G'
are homeomorphic, then their complements ~n \ G and ~n \ G' have the
same number N ~ 00 of connected components.

Proof. Let j : G c ~n be the canonical inclusion. Since j is a proper


mapping, it extends to the one-point compactifications G* of G and §n of
~n, by mapping 00 f---t 00. Similarly, j' : G' c ~n extends to G'* -+ §n. Of
course, G* and G'* are homeomorphic as G and G' are. Thus, we have two
homeomorphic compact subsets G* and G'* of the sphere §n, and by the
preceding proposition, they disconnect §n in the same number of connected
components. Note here that none of those components contains 00; hence
they pull back to ~n without surprise, and the result is proved. 0

For instance, we can improve the statement of IV.7.2(1), p. 180, to the


following: a closed set G c ~n that is homeomorphic to a closed differen-
tiable hypersurface disconnects ~n.
We close this section with a new proof of the Invariance of Domain
Theorem:
Proposition 7.5. Let W c ~n be an open set. Every locally injective
continuous mapping f : W -+ ~n is open.
182 IV. Degree theory in Euclidean spaces

Proof. Pick any a E W. Consider an open ball B centered at a of radius


E> 0 and the compact sets
K = {llx - all ~ c} and L= {llx - all = c};

by hypothesis, for E small enough, f is defined and injective on K. Then,


the restrictions flK and flL are injective closed mappings (by compactness
of their domains of definition), hence homeomorphisms onto their images
f(K) and f(L). Consequently, jRn\f(K) is connected (and unbounded) and
jRn \ f(L) splits into two connected components Do, Dl, with Dl bounded.
It follows that jRn \ f (K) c Do. Hence

Dl c f(K) = f(B) U f(L),


and we conclude that Dl C f(B). But f(B) is a connected subset of
jRn \ f(L), so it cannot be bigger than D 1 . Thus f(B) = Dl is an open
set. This shows that f maps all small enough open balls centered at a onto
open sets. Since a is arbitrary, f is open. 0

Exercises and problems


Number 1. Prove the following general facts for a compact set K in Rn.
(1) The complement R n \ K has exactly one unbounded connected component.
(2) The set K contains the boundaries of all connected components of R n \ K.
(3) If R n \ K is not connected, but R n \ A is connected for every closed subset A s;: K,
then K is the boundary of all connected components of R n \ K.
Number 2. Exhibit two homeomorphic compact subsets of R3 whose complements are
not homeomorphic.
Number 3. Let C and C' be closed sets of RP and Rq, respectively. Show that if C and
C' are homeomorphic, then RP x Rq \ ex {O} and RP x Rq \ {O} X C' are homeomorphic.

Number 4. Let "(, "(' : [0, 1] ~ D2 be two injective continuous mappings with

"((0) = (1,0), "((1) = (-1,0); "('(D) = (0,1), "('(I) = (0,-1).


Prove that they meet at some point.
-2 -2
Number 5. Let C be a closed subset of the closed disk D . If C and D \ C are both
-2 -2
connected, then en aD and aD \ C are both connected.
Number 6. Let M be a boundaryless connected smooth manifold, dim(M) 2 2.
(1) Find some smooth hypersurface HeM, diffeomorphic to a sphere, such that
M \ H has two connected components.
(2) Show that there may be two hypersurfaces Hl, H2 eM, both diffeomorphic to
a sphere, such that M \ Hl has two connected components and M \ H2 is connected.
Chapter V

The Hopf Theorems


In this chapter we formulate and prove the theorems that fully complete degree
theory: the degree is the only homotopy invariant for spheres. This must be for-
mulated for both the Brouwer-Kronecker and the Euclidean degrees, although in
essence the two cases are the same. These are the Hopf Theorems. As these the-
orems refer to mappings into spheres, we start the chapter by constructing in §1
various basic examples of them. Then in §2 we prove the first Hopf Theorem, that
two mappings into a sphere of the same (Brouwer-Kronecker) degree are homo-
topic; hence, in particular, the cohomotopy group 7rm(M) of any m-manifold and
the homotopy group 7rm(sm) of the m-sphere are Z. In §3 we deduce the same re-
sult for the Euclidean degree, which has the virtue of determining the cohomotopy
groups 7rm(x) for some more exotic spaces X of dimension m, like the topological
circle. These Hopf Theorems require the use of diffeotopies. Next, in §4 we come
back to the Hopf invariant and define the Hopf fibrations to study some homotopy
groups 7rk (sm) for k > m. Then we turn to the theory of tangent vector fields, and
in §5 we discuss the essential notion of index of a tangent vector field at an isolated
zero, using both the Brouwer-Kronecker and the Euclidean degrees. This notion
is best illustrated by the gradients of Morse functions, which we describe in §6.
Finally, we conclude the chapter and the book by presenting in §7 another deep
theorem also named after Hopf, the Poincare-Hopf Index Theorem, that computes
the index of a tangent vector field. We also see how the Gauss-Bonnet Formula
follows from this.

1. Mappings into spheres


In this section we discuss several methods for constructing smooth map-
pings into spheres, or to modify one already given.
To start with, we exhibit simple examples of mappings of arbitrary
degree.
Example 1.1. Let §m C jRm+l be the standard unit sphere. Then, for
every integer d there are mappings fd : §m -+ §m of degree d.
Since all constant mappings have degree 0, we assume d =I O.
-183
184 v. The HopE Theorems

Case d ~ 1. We will generalize to dimension m the intuitive idea of winding


a circle k times around another (III.1.6(3), p. 100). Any point x E §m can
be parametrized as

x = (Xl,X2,X') = (pcosO,psinO,x') where p = .j1-lix'11 2 .


Then, we define the winding by

9d(X) = (pcos(d·O), psin(d·O), x').


To eliminate the parameter 0, recall that

cos(d·O) = Pd(cosO,sinO), sin(d·O) = Qd(cosO,sinO),

for suitable homogeneous polynomials Pd and Qd of degree d. We have

9d(X) = (PPd(~Xl,~X2),pQdGXl,~X2)'X')
= (pLl Pd(Xl, X2), i-I Qd(Xl, X2), x').
This shows that the map 9d is differentiable except when lix'il = 1. To
amend this, we use the deformation ji of p defined by

for a smooth function A(t) which is = t for t ~ ! and < t for t > !, as in
the figure below:

Thus, we have the smooth mapping

Yd(X) = (jid~l Pd(Xl, X2), jiLl Qd(Xl, X2), x').


For Ilx'll ~ !
this coincides with 9d(X) E §m, but otherwise, we must
normalize to get fd = Yd/IiYdll : §n -+ §n. We claim that deg(fd) = d.
Indeed, pick a = (1,0, ... ,0) E §m. Suppose fd(X) = a. Then x' = 0
and p = 1, so that fd(X) = 9d(X) = (cos(d·O), sin(d·O), 0), and we obtain
k
where Ok = 27rd ,0 ~ k < d.
1. Mappings into spheres 185

Near such a point we can use the parametrization x = (p cos 0, p sin 0, x'),
which is compatible with the standard orientation of the sphere (straight-
forward computation). The local expression of our mapping is then

fd == 9d : (0, x') t-+ (d·O, x').

This is clearly a diffeomorphism that preserves orientation, which shows


that a is a regular value of fd, and
d
deg(Jd) = L signx(Jd) = L(+l) = d.
xEfil(a) k=1

One can use the same method for d negative, just noting that one winds
around clockwise, instead of counterclockwise. However, here we have the
following alternative argument.
Case d = -1. As we saw in 111.1.6(2), p. 100, the symmetry

has degree deg(J-l) = -1.


Case d < 0. Define fd = f-l 0 f-d, so that by 111.1.7, p. 100,

deg(Jd) = deg(J-l) . deg(J_d) = (-1)( -d) = d. o

We can extend the preceding result to arbitrary compact manifolds


(compactness is required because the target §m is compact):
Proposition 1.2. Let M be a compact, oriented, boundaryless, differen-
tiable manifold of dimension m. Then for every integer d there is a differ-
entiable mapping hd : M -t §m of degree d.

Proof. First we remark that it is enough to find h : M -t §m of degree


±1. Indeed, given such an h, we can use the mappings f±d of the preceding
example to define hd = f±dOh, which has degree d (111.1.7, p. 100). In fact,
let us define a differentiable mapping h : M -t §m such that the north pole
(0, ... ,0,1) E §m is a regular value with a single preimage.
Let U be an open subset of M diffeomorphic to ~m, and consider a
smooth function
!,
= ~(t)
for t ::;
for t ~ 1.
G
186 v. The HopE Theorems

We identify U == JR.ffi to define h : M -+ §ffi by

(O, ... ,0, -1) the south pole, for x ct. U,


{
h(x)= ( 2T(X)X T(x)2 -llxIl2 )
where T(x)=p,(llxI1 2 ), for x E U.
T(x)2 + Ilx11 ' T(X)2
2 + IIxll 2
This mapping is well defined because for x E U with IIxll ~ 1 we have
°
T(X) = and h(x) = (0, ... ,0, -1). On the other hand, using the equations
of the stereographic projection 7r from the south pole (11.2.4, p. 55), for
x E U with T(X) =f:: 0, we get

h(x) = 7r-I(T~X)).
Thus, h is the diffeomorphism 7r- 1 for T(X) = 1, that is, for Ilxll ::; On !.
°
the other hand, x = is the only preimage of (0, ... ,0,1). Indeed, suppose
(0, ... ,0, 1) = h(x). Then T(X) =f:: 0, and

(0, ... ,0,1)=h(x)=7r -l( T(X)


x ) j

hence x = 0. We are done. D

Remark 1.3. The existence of non-homotopic maps M -+ N between


two manifolds depends on the nature of the given manifolds. In general
we cannot expect the richness shown in V.1.2 above. To illustrate the
opposite situation we recall that all continuous mappings §2 -+ §l X §l are
null-homotopic, hence have degree zero.
Indeed, the reason is that any continuous mapping f : §2 -+ §l X §l can
be lifted to 1:§2 -+ JR. x JR., through the universal covering

7r: JR. x JR. -+ §l X §l : (a,{3) H (cosa,sinajcos{3,sin{3),


-
so that f :::: 7r 0 f. Then the homotopy Ht(x)
-= (1-
- t)f(x) goes down to
Ht = 7r 0 H t with Ho = f and HI constant. D

Now we want to modify mappings already given. The modification must


keep the homotopy type. For mappings into spheres this comes easily, as
follows:
Lemma 1.4. Let M be a possibly non-compact manifold. Let f, g : M -+
§ffi be two mappings that do not have antipodal images: f(x) =f:: -g(x) for
all x E M. Then
H x _ tg(x) + (1 - t)f(x)
t( ) - Iltg(x) + (1 - t)f(x)11
1. Mappings into spheres 187

is a well-defined homotopy with Ho = f, HI = g.


In particular, if f is not surjective, it is null-homotopic.

Proof. Let x E M. Then p = tg(x) + (1 - t)f(x) is a point in the segment


joining f(x) and g(x). As both points belong to the sphere, p = 0 only if
f(x) and g(x) are antipodal, which is excluded by hypothesis. Thus Ht is
indeed well defined.
Finally, if there is a p ¢. f(M), f is homotopic to the constant mapping
9 == -po D

An immediate application of the lemma is that every differentiable map-


ping M -7 §m with dim(M) < m is null-homotopic. Indeed, with that
hypothesis, no smooth mapping M -7 §m is surjective, by the Easy Sard
Theorem.
Now that we have this homotopy lemma, we will further explore the
idea in the proof of V.1.2, p. 185:

(1.5) Antipodal extension, linearization, and normalization. Let


M be an oriented, boundaryless, differentiable manifold of dimension m,
possibly non-compact.
(1) Suppose we fix some parametrization ~m == U c M and a smooth
mapping h : ~m -7 ~m such that h-I(O) = {c}. Then we use the stereo-
graphic projection 7r : §m -7 ~m from the south pole as = (0, ... ,0, -1) E
§m to identify ~m with the open set V = §m \ {as} C §m and E ~m
with the north pole aN = (0, ... ,0,1) E §m. Thus, h == 7r- 0 h is seen as
I
°
a mapping into ~m == V C §m :
M §m
U C ~ O==aN U
U == IRm ~ IRm == V
What we want is to extend h : U -7 V from a neighborhood of c to
Ii : M -7 §m by sending everything off another neighborhood of c to the
south pole.
For such an extension we modify h by a trick analogous to the one used
in the proof mentioned above:
(O, ... ,0, -1) for x ¢. U,
{
Ii(x) = 2T(X)h{x) T{x)2 - Ilh{x)1I2
for x E U,
(T(X)2 + IIh(x) 112 'T{x)2 + IIh{x) 112 )
188 V. The HopE Theorems

with a smooth bump function

I for Ilx - cll :S~,


T(X) = J.L(llx - c1l 2) = {0
for Ilx - cll ~ 1.
This defines well a smooth mapping, because Ilh(x)11 f:. 0 on a neighborhood
of IIx - cll ~ 1, and Ii(x) = (0, ... ,0, -1) for IIx - cll ~ 1. After some
straightforward computation, for x E U with T(X) f:. 0 we find

Ii(x) = 7r-1 (h(X));


T(X)

hence h == 7r- 1 oh on the ball ofradius~. Note also that Ii(x) = (0, ... ,0,1)
if and only if x E U and h(x) = 0, if and only if x = c. We will call this
construction the antipodal extension.
In view of the homotopy lemma, V.1.4, p. 186, we are also interested in
possible antipodal images of Ii and h in §m. Suppose this is the case for
some point x E U. Then

( 2T(X)h(x) T(x)2 -lIh(x)112 ) ( 2h(x) 1 - Ilh(x)1I2 )


T(x)2 + IIh(x)112 'T(x)2 + Ilh(x)112 = - 1 + Ilh(x)112 ' 1 + Ilh(x)112 .
Looking at the last component of these two points, we find that T(X) =
Ilh(x)112 and h(x) =1= O. Then, looking at the other components, we get

1 + Ilh(x)112'
which is impossible. Hence we conclude that the mapping h and its exten-
sion Ii have no antipodal images.
(2) Next, suppose h is a diffeomorphism, and denote L = dch, which is
a linear isomorphism. Consider the homotopy of IV.3.1, p. 153, which here
simplifies to
h(tx + (1 - t)c) £ ...J. 0
H t (X ) -_ { or t r ,
t
L(x - c) for t = O.
Clearly Ht(x) = 0 if and only if x = c, and we can apply the same antipodal
extension to H t replacing h by H t all through the construction. We get a
homotopy fIt such that fIl = Ii and 7r 0 fIo == L - L(c) near c. We have
thus linearized the mapping Ii.
(3) Now, the determinant of the linear isomorphism L can be either
positive or negative. In the first case, let A denote the identity, and in the
1. Mappings into spheres 189

second, the symmetry with respect to the first variable in jRm. In any case,
the determinants of L and A have the same sign. Now, m x m matrices are
points in jRmxm, and the closed hypersurface det = 0 decomposes jRmxm
into two connected components: det > 0 and det < O. Since L and A are
in the same one of those two components, there is a continuous mapping
tHAt with det(At) =I- 0 for all t, Ao = L and Al = A. Again we apply the
antipodal extension to B t = At-At(c) and find a homotopy 13t . We remark
here that the construction is possible because At is bijective (det(At) =I- 0):
this guarantees that B t (x) = At (x - c) = 0 if and only if x = c. This
normalization shows that the mapping h is homotopic to 9 = 131 , which is
the antipodal extension of A - A(c).

The construction we have just described is the method for classifying


smooth mappings f : M -+ §m by homotopy. Indeed, suppose the north
pole a E §m is a regular value of f with finitely many preimages Cl, ... , Cr.
As usual, we find local coordinates around each Ci, on whose domains Ui the
restrictions flui are diffeomorphisms. Choosing the Ui'S disjoint and small
enough, we can glue the antipodal extensions of the restrictions flui into
a smooth mapping J: M -+ sm. By (1) above, f and J have no antipodal
images, and by V.1.4, p. 186, f and J are homotopic. This mapping J
depends only on the preimages of the regular value and the signs of f at
them: exactly what we use to compute the degree.
Actually, this is almost a proof of the Hopf Theorem, and what remains
is quite evident: (i) to move the preimages freely and (ii) to dispose of
those pairs whose signs cancel each other. We devote the following section
to settling these two matters, by means of diffeotopies (II.6, p. 80).

Exercises and problems


Number 1. Points (x, y, t) in the sphere §2 C lR? == c x lR can be written as pairs (z, t),
z = x +iy. For every integer k > 0 define a continuous mapping f : §2 -t §2 by a formula
f(z, t) = (Zk, th(t)) involving a suitable function h > O. Compute the degree of f.
Number 2. It is a fact that all smooth mappings Cpl x Cpl -t CP2 have even degree.
Mappings of every possible degree can be obtained as follows.
(1) Use the Segre embedding

Cpl x Cpl -t Cp3: (xo: Xl;YO: Yl) f-t (XOyo: XOYI : XIYO: XIYl)

and a conic projection in Cpa to produce a smooth mapping Cpl x Cpl -t CP2 of degree
2.
(2) Compose the mapping in (1) with mappings Cpl x Cpl -t Cpl XCpl of arbitrary
degree.
190 V. The HopE Theorems

Number 3. Let f : §m -+ §m be a continuous mapping. Show that:


(1) If deg(f) =1= +1, there is an x E §m with f(x) = -x.
(2) If deg(f) =1= (_1)m+1, there is an x E §m with f(x) = x.
(3) If f is null-homotopic, then it fixes a point and sends another to its antipode.

Number 4. Let f, 9 : §2n -+ §2n be two continuous mappings. Prove the following
assertions:
(1) At least one of the mappings f, g, or go f has a fixed point.
(2) If f has no fixed point, there are x, y E §2n such that f(x) = y and f(y) = x.
(3) If f has no fixed point, then it sends some point to its antipode.

Number 5. Let r : §m -+ §m be a continuous involution (=1= Idsm). Show that r sends


some point to its antipode by way of contradiction, as follows:
(1) Prove that, otherwise,

f §m §m x+r(x)
: -+ :XH IIx+r(x)11
is a well-defined continuous map of degree 1, hence surjective.
(2) Let F C §m denote the set of fixed points of r, and write V = §m \ F, U =
f-l(V) C §m. Show that flu: U -+ V is a well-defined proper map of the same degree
as f.
(3) Use the identification x == r(x) to define a surjective local difeomorphism 1T :
U -+ M onto an abstract smooth manifold M (just as x == -x gives M = um).
(4) Show that flu factorizes through M, and use that factorization to conclude that
f has even degree.
Number 6. Let G be a non-trivial group of homeomorphisms acting on a sphere of
even dimension. Suppose that no h E G distinct from the identity has fixed points. Then
G==Z2.

Number 7. Let f, 9 : §m -+ §m be two continuous mappings which are never perpen-


°
dicular (that is, (f(x),g(x») =1= for all x E §m). Then deg(f) = ±deg(g).
Number 8. Let P, Q denote two complex polynomials of degrees p and q, respectively,
and consider the rational function F = P / Q : C == ]R2 -+ C == ]R2. This function is
defined except at finitely many poles Cl, ... ,Cr : the zeros of Q which do not cancel with
those of P. Show that via the sterographic projection from the north pole aN = (0,0,1),
F extends to a smooth mapping f : §2 -+ §2 by setting

Compute the degree of f.


Number 9. Let M C ]RP be a connected, boundaryless, smooth manifold of codimension
m. Show that there is a differentiable mapping f : ]RP -+ §m such that the north pole
aN = (0, ... ,0, 1) E §m is a regular value of f and M = rl(aN).
2. The HopE Theorem: Brouwer-Kronecker degree 191

2. The Hopf Theorem: Brouwer-Kronecker


degree
Here we will show that an example like III.3.5, p. 112, cannot occur for
mappings into spheres.

Theorem 2.1 (Hopf Theorem). Let M be a connected, compact, oriented,


boundaryless, differentiable manifold of dimension m. Two continuous
mappings M -+ §m with the same Brouwer-Kronecker degree are homo-
topic.

Proof. We will prove the theorem by showing that all mappings f : M -+


§m of a given degree have a common normal form, in the sense described
in the previous section. Suppose deg(f) = d ~ 0 (resp., d ::; 0) and fix a
coordinate domain U == ~m, whose local coordinates preserve orientation,
and any d points CI, ... , Cd E U (none if d = 0). We are to built up a
mapping homotopic to f that solely depends on these points and d.
We can assume f is differentiable and choose a regular value a E §m,
which by a diffeotopy (11.6.5, p. 82) of the sphere we move to the north
pole. Then a has 2r + d inverse images, so that
(i) at d + r of the images the sign of f is positive (resp., negative),
(ii) at r of the images the sign of f is negative (resp., positive).
Then, choose 2r + d points PI, qI, ... ,Pr. qr, CI, ... ,Cd E U, and use a dif-
feotopy to move the points in f- l (a) as follows (we do nothing if there are
no points to move):
(1) r of them at which f is negative (resp., positive) to ql,· .. ,qr,
(2) some r at which f is positive (resp., negative) to PI,··· ,Pr,
(3) the remaining d of them at which f is positive (resp., negative) to
cI,···, Cd·
Using the chosen coordinates in U, we identify U == ~m, and our points
qk,Pk, Ci, are points in ~m. Then we use a diffeotopy of ~m which is the
identity off a big enough ball (II.6.6, p. 83) to move those points so that
(1) qk = (-3k, 0, ... ,0),
(2) Pk = (3k, 0, ... ,0),
(3) Ci = (3(s + i), 0, ... ,0), for some fixed s ~ r.
192 V. The HopE Theorems

Now, using the antipodal extension, linearization, and normalization (V.1.5,


p. 187), we find a mapping J homotopic to j, which is == -a off the open
balls of radius 1 around the points qk,Pk, Ci, and inside those balls it is,
respectively, the antipodal extension of
(1) the symmetry: (Xl, y) f---t (-Xl, y) + qk,
(2) the translation: (XI,Y) f---t (XI,Y) - Pk, and
(3) the translation: (Xl. y) f---t (Xl, y) - Ci·
What is important here is that J(-XI,Y) = J(XI,Y) for IXII ~ 3r+ 1. This
implies that the following homotopy is well defined for ~ t ~ 3r + 1: °
H t ( Xl, Y) -_ {J(XI,Y) for IXII ~ t,
-
j(t, y) for IXII ~ t.

Note that this homotopy on]Rm == U c M extends to M, because H t == -a


off -3r - 1 ~ Xl ~ 3(8 + d) + 1, IIYII ~ 1. We have Ho = J, and for
t = 3r + 1:

_ {J( x) on the balls centered at the points Ci,


H t (X ) -
-a off those balls.

This 9 = H t is the mapping we seek. Of course, it is homotopic to J,


hence to j, and, in addition, the north pole a is a regular value of g,
g-l(a) = {CI, ... ,Cd}, and 9 is the antipodal extension of the translation
(Xl, y) f---t (Xl, y) - Ci on the unit ball around each Ci. Note that if d = 0,
this is the constant mapping == -a.
Clearly, this construction can be carried along simultaneously for two
mappings of degree d taking enough points Pk, qk, and a common 8, and
the final form is the same for both. 0
2. The Hopf Theorem: Brouwer-Kronecker degree 193

Corollary 2.2. Let M be a connected, compact, oriented, boundary less,


differentiable manifold of dimension m. Then

Proof. This follows immediately from V.2.I, p. 191, and V.1.2, p. 185, be-
cause 7rm (M) = [M, §m]. D

Remarks 2.3. (1) The condition that M is boundaryless is essential to


have different homotopy classes. In fact, if M has a boundary 8M f:. 0,
then all mappings M -+ §m are null-homotopic.
We only present a quick sketch of the argument. First one glues along
their boundaries M and a copy - M with the opposite orientation, to get a
new oriented, compact, boundaryless manifold M* = M u - M of the same
dimension (some smoothing is required around 8M == 8( -M) by means of
a bump function). Then every continuous mapping f : M -+ §m extends
in an obvious unique way to f* : M* -+ §m, and this f* has degree zero
(pick a close smooth approximation 9 of f* and look at a regular value of 9
off g(8M)). By the Hopf Theorem, the mapping f* is null-homotopic, and
so is its restriction f = f*IM.
(2) After the above example, it is natural to ask for a degree for man-
ifolds with boundary. This can indeed be done, using the ideas behind
Euclidean degree. If f : M -+ N is a proper mapping from an (m + 1)-
manifold with boundary into a connected, boundaryless, (m + I)-manifold,
there is a degree d(f, M, a) for a E N \ f(8M). In this setting homotopies
do not touch the boundary and the situation in (1) does not occur.
(3) In case M is non-compact, there is no proper mapping M -+ §m,
so that there is no notion of degree for a mapping f : M -+ §m. But still,
one can ask about the homotopy type of f. The dramatic answer to this
question is that f is always null-homotopic. Actually, the argument is a
kind of infinite degree computation. We only sketch the idea.
First, replace f by some close smooth mapping, which will be homo-
topic. Then, pick a regular value a E §m, and move it to the north pole.
Replace f with a homotopic mapping which is == -a off a compact neigh-
borhood of f-l(a) (the antipodal extension is still posible because f-l(a)
is discrete). Pick a countable discrete family of coordinate domains in M
far from that neighborhood, and reversing the method used in the proof
of the Hopf Theorem to cancel points with opposite sign, create a pair of
194 V. The HopE Theorems

them on each such domain. From this we are sure that f- 1 (a) consists of
an infinite sequence (Pk) of points at which the sign of f is positive and
another sequence (qk) at which the sign is negative. Then we can move
each pair Pk, qk into a coordinate domain far from the other points, using a
diffeotopy that does not move anything else. Thus they cancel each other
separately as in the proof of the Hopf Theorem, and in the end, f-l(a) is
empty. Thus, f is not surjective, hence null-homotopic.
(4) For non-orientable manifolds, the mod 2 Brouwer-Kronecker degree
(III.6.1, p. 124) provides the following parallel version of the above Hopf
Theorem, V.2.1, p. 191:
Let M be a connected, compact, non-orientable, boundaryless, dif-
ferentiable manifold of dimension m. Two continuous mappings
M --t §m with the same mod 2 Brouwer-Kronecker degree are ho-
motopic.
In particular, we see that for such an M, 7rm(M) = Z2.
The proof is no surprise: one can cancel pairs of points p, q in the inverse
image of any regular value. Indeed, after some diffeotopy, non-orient ability
gives two parametrizations <p, 'ljJ : ]Rm --t U, V with p, q E Un V that induce
different orientations at p, but the same at q. Hence the localization of the
given mapping via one of those parametrizations must preserve orientation
at p and reverse orientation at q (or the opposite), and then we eliminate
the pair as in the orient able case. 0

From the Hopf Theorem we easily deduce for spheres the converse to
the Boundary Theorem, III.3.3, p. 111:
Proposition 2.4. Let X be an oriented differentiable Cr manifold of di-
mension m + 1 with connected boundary ax = M, and let f : M --t §m be
a proper (resp., proper Cr ) mapping of degree O. Then f has a continuous
(resp., Cr ) extension J : X --t §m.

Proof. As f is proper, M is compact. By V.2.1, p. 191, there is a homotopy


H t : M --t §m with Ho == a E §m and HI = f. Then, the set Z =
({O} x X) U ([0, 1] x M) is closed in [0,1] x X, and we have the continuous
mapping

F:Z--t
§m ( )
: t,x H
{a for t = 0,
Ht(x) forti-O.
Now, by the Tietze Extension Theorem, this extends to a continuous map-
ping F : [0,1] x X --t ]Rm+1. Consider the set A = F- 1 (]Rm+l \ {O}), which
2. The HopE Theorem: Brouwer-Kronecker degree 195

is an open neighborhood of Z in [0,1] x X. In particular, there is an open


set U C X such that U ~ M and [0, 1] x U c A, and we choose a bump
°
function 0 which is == off U and == 1 on M. We define a continuous
mapping J : X -+ §m by

- F(O(x), x)
f(x) = IIF(O(x), x)II'

This is well defined, because F(O(x),x)


x¢. U and O(x) = 0, so that
= °implies (O(x), x) ¢. A; hence

0= F(O(x), x) = F(O, x) = a E §m,


a contradiction. Furthermore J is an extension of f, since O(x) = 1 on M,
so that JIM = HI = f.
We have thus proved the assertion for proper mappings. For proper Cr
mappings the argument is the same using the Cr Tietze Extension Theorem,
11.1.5, p. 52. The only requirement is to change the first definition:

F: Z -+
§m ( )
: t, x t---+
{aHTf(t) (x) for t
for t
= 0,
i= 0,
° k
where'T/ is a smooth bump function == for t ~ and == t for t 2 ~. This
guarantees F is Cr , and the proof follows readily. D

In fact the above proof is an adapted version of the general argument


used to extend homotopies from closed subsets of a metric space to the
whole space (the Borsuk Theorem). This then implies that the property
that a continuous mapping on such a closed subset extends to the whole
space only depends on the homotopy type of the mapping.

Exercises and problems

Number 1. Let f, 9 : §m --+ §m be two continuous mappings. Prove that fog and go f
are homotopic.
Number 2. Let f, 9 : §l --+ §1 be smooth mappings. Then the (complex) multiplication
f . 9 : §1 --+ §1 is well defined, and
deg(f . g) = deg(f) + deg(g).
Prove this directly from the Hopf Theorem and compare with Problem Number 8 of
111.3. Deduce that if a complex polynomial P(z) has m roots (counted with multiplicities)
inside the circle §l C ]R2 == C and none on it, then the mapping f : §l --+ §l : z I-t l~~:ll
has degree m.
196 V. The HopE Theorems

Number 3. Use the preceding formula deg(J . g) = deg(J) + deg(g) to review Problem
Number 4 of IlIA, noting that for the 9 and h there, h = Id 'g.
Number 4. Let X be a space with involution for which s(X) = s'(X) = m < 00
(Problem Number 6 of IV.5). Then 7rm (X) contains Z.
Number 5. Let M and N be smooth manifolds, N boundaryless and connected, and
fix a point c EN. For every proper smooth mapping f : M --+ N such that f(8M) = {c}
define
deg(J) = 'E signx(J),
xEf-l(a)

where a t- c is a regular value of f. Show that this degree is well defined and that it is
invariant by homotopies H t such that H t (8M) = {c}.
Number 6. Two spaces X and Y have the same homotopy type if there are continuous
mappings f : X --+ Y and 9 : Y --+ X such that go f and fog are homotopic to Idx and
Idy. Prove the following:
(1) Two spheres of different dimensions do not have the same homotopy type.
(2) A sphere does not have the same homotopy type as the product of two others.
Number 7. Let N C Me RP be differentiable manifolds, N closed in M. Let X be a
topological space and let H : [0, 1] x N --+ X be a continuous homotopy. Suppose that
f = H 0 : N --+ X extends to a continuous mapping ! : M --+ X. Then H extends to a
continuous homotopy H: [0,1] x M --+ X such that! = Ho.
Number 8. Let X be a metric space, C C X a closed subset, and H : [0,1] x C --+ M
a continuous mapping into a differentiable manifold M. Suppose that f = Ho : C --+ M
extends to a continuous mapping! : X --+ M. Then H extends to a continuous homotopy
H: [0,1] x X --+ M such that! = Ho.

3. The Hopf Theorem: Euclidean degree


In this section we prove the suitable version of the Hopf Theorem, V.2.1,
p. 191, in the Euclidean degree setting:

Theorem 3.1 (Hopf Theorem). Let X C ~m+1 be a compact set that


disconnects ~m+1 into two connected components, DI bounded and Do un-
bounded, whose topological boundary is X. Let j, g : X --+ ~m+1 \ {O} be
two continuous mappings with the same winding number: w(j,O) = w(g, 0).
Then j and g are homotopic.

Proof. The proof is a recollection of various results already at hand. After


a translation, 0 E D 1 , and we choose an open ball B of radius p containing
D 1 , whose boundary (the sphere of radius p) we denote by M = B \ B.
Now recall that

w(j,O) = d(J, Db 0) and w(g,O) = d(g, D 1 , 0),


3. Tile HopE Tileorem: Euclidean degree 197

where J, 9 are arbitrary continuous extensions of f, 9 to Dl = D1 U X,


which we can further extend to B, without zeros in M (Tietze Extension
Theorem). By the same definition of degree (IV.2.1, p. 145, and IV.2.2, p.
146), we can assume that J and 9 are smooth on B and that 0 is a regular
value for both on B. Actually, J and 9 are polynomials; hence they are
defined on the whole Euclidean space ~m+1.
Let us now concentrate on f. The discrete set J- 1 (0) n B is finite,
and its points split into those in D1 and those in Do n B (note that 0 ~
f(X) U f(M)): say C1, ••. , Cr E DI, bI, ... , bs E Do n B. Now pick an open
ball B' C D1 centered at 0 E D 1. Since D1 is connected, we can move the
cj's into B' by a diffeotopy of ~m+1 that fixes X and all other zeros of J;
then, we pick a ball B" :J B, so that J has no zeros in B" \ B, and since
D1 n B" is connected, we move the bi'S into B" \ B by a diffeotopy of ~m+1
that fixes X and all other zeros of J (II.6.6, p. 83).

o
o
o

- 1 ( .0

~J(X)

We have moved the points so that 0 has the same preimages in Band D1;
hence d(J, D1 , 0) = d(J, B, 0). Furthermore the mapping

1: u = B \ B' -+ §m : x H ~(x)
IIf(x)11
is well defined. Now, by IV.4.5, p. 157, the Euclidean degree d(J, B, 0) is
the Brouwer-Kronecker degree of the smooth mapping M; that is,11
198 v. The HopE Theorems

The same construction for 9 gives a smooth mapping

'9: u = B \ B' -+ §m : x I-t g(x)


Ilg(x)11
with
d(g, Db 0) = deg('9IM).

Con~quently deg(fiM) = deg('9IM), and by the Hopf Theorem, V.2.1,


p. 191, flM and '9IM are homotopic. It follows immediately that the two
mappings

f* : X -+ §m : x I-t f(px/llxID and g*: X -+ §m : x I-t '9(px/llxll)

are also homotopic or we can simply say that f* : X -+ jRm+l \ {O} and
g* : X -+ jRm+l \ {O} are homotopic. Thus, it remains to show that f* and
g* are, respectively, homotopic to f and g.

We write the proof for f in two steps. First, since 1is defined on B \ B',
we can consider the homotopy

Ht(x) = f(tx + (1 - t)px/llxll) i= 0, x E X

(every segment joining x E X to px/llxll EM is fully contained in B \ B').


Clearly Ho = f* and HI = fix.
Second, we show that fiX is homotopic to f. To that end, consider the
mapping
Ft(x) = /(x) _
(1 - t) + tllf(x)11
If the denominator vanishes at some x E X, then t = 1 and /(x) = 0,
which is impossible. Also, if Ft(x) = 0 for some x E X, then /(x) = 0,
again impossible. Hence, Ft is a well-defined homotopy with Fo = fix = f
and Fl = fix.
For 9 the argument is the same, and the proof of the theorem is thus
finished. D

As in the preceding section, this completely describes the cohomotopy


group:
Corollary 3.2. Let X C jRm+l be a compact set that disconnects jRm+l
into two connected components whose topological boundary is X. Then
3. The HopE Theorem: Euclidean degree 199

Proof. The group 7l'm(x) classifies by homotopy the mappings X -+ §m.


Using the inclusion §m C lR. m+1 and its canonical retraction lR.m+1 \ {O} -+
§m : x t-+ x/llxll, one sees immediately that 7l'm(x) = [X, lR. m+1 \ {O}l.
Then, by the latter Hopf Theorem, V.3.1, p. 196, the winding number gives
an inclusion [X, lR.m+1 \ {O}l c Z, and to conclude, it is enough to exhibit
mappings f : X -+ lR. m+ 1 \ {O} of arbitrary degree. This follows easily by
the technique used in the preceding proof.
Let Dl be the bounded component of lR. m+1 \ X, and pick any big ball
B ::J X U D 1 • Let M denote the sphere that bounds B and choose a smooth
mapping h : M -+ §m C lR. m+1 of Brouwer-Kronecker degree deg(h) = d
(V.1.1, p. 183); by IV.4.6, p. 158,

d = deg(h) = w(h, 0).


Let h : B -+ lR. m+1 be a smooth extension of h. Then d = w(h,O) =
d(h, B, 0) is computed through the preimages of any regular value a of h.
Now, those preimages off Dl can be moved inside by a diffeotopy that leaves
kf invariant. Consequently, we can assume all preimages are in D 1 , so that

d = d(h, B, 0) = d(h, D 1 , 0) = w(hlx, 0).

We are done. D

Also, we can deduce the converse of the Boundary Theorem for winding
numbers (IV.4.4, p. 157):

Proposition 3.3. Let f : X -+ lR.m+1 \ {O} be a continuous mapping


with winding number w(j,O) = O. Then f has a continuous extension
J: D -+ lR.m+1 \ {O}.
The proof is a copy of that of V.2.4, p. 194, using V.3.1, p. 196, instead
of V.2.1, p. 191. We leave it to the reader.

Exercises and problems

Number 1. We have shown, by a combination of various results, that the Hopf Theorem
for the Euclidean degree (V.3.1, p. 196) follows from the Hopf Theorem for the Brouwer-
Kronecker degree (V.2.1, p. 191). Show that to the contrary, the converse implication is
quite immediate for hypersurfaces.

Number 2. Let h : IR -+ IR be a periodic continuous mapping of the form

hex + 2rr) = hex) + 2krr, for a fixed integer k.


200 V. The HopE Theorems

Compute the winding number w(j,O) of the continuous mapping I : Sl --t IR2 \ {O}
induced by h.
Number 3. Let I : sm --t IRm+1 \ {O} be a continuous mapping such that

Then identify sm-l == sm-l X {O} and IR m == IRm x {O}, so that I restricts to a mapping
9 = Ilsm-1 : sm-l --t IR m\ {O}. Prove the following:
(1) w(g,O) = w(j, 0).
(2) The suspension of glllgil (Problem Number 4 of IV.6) is homotopic to 1/11111.
Number 4. Let X C R2 be the curve of Problem Number 50fIV.I. Show that although
X is not homeomorphic to a circle, it verifies the hypotheses of V.3.1, p. 196, and find
simple representatives for all homotopy classes in [X,IR 2 \ {O}l.

4. The Hopf fibration


The Hopf Theorems in the preceding sections show how the degree fully
determines the cohomotopy groups [M, §m] = Z when dim(M) = m. We
also know that [M, §m] = {O} when dim(M) < m (V.1.4, p. 186). Thus
the problem is the case dim(M) > m, in particular for spheres M = §k.
However, the degree can still be used to obtain important information con-
cerning mappings §2m-1 --+ sm. The tool is the Hopf invariant introduced
in III.5.1, p. 119. The additional key property of this invariant that we give
here follows.
Proposition 4.1. Let 9 : §2m-1 --+ §2m-1 and f : §2m-1 --+ §m be smooth
mappings. Then
H(f 0 g) = H(f) deg(g).

Proof. We can assume m even, because otherwise H == O. Since the Hopf


invariant and the degree depend only on the homotopy class, we can at
any moment change every mapping by any other one homotopic to it. This
said, we distinguish several cases.
Case I: deg(g) = O. Then 9 is null-homotopic, and we can replace it with a
constant function, so that fog is also constant, and the assertion is trivial.
Case II: deg(g) = 1. Then 9 is homotopic to the identity, and for the
identity the result is evident.
Case III: deg(g) = -1. We can assume that 9 is the linear symmetry with
respect to the hyperplane Xl = 0, which is compatible with the stereo-
graphic projection from the south pole, and then 9 = g-l. Pick regular
4. Tbe Hopi fibration 201

values a, b to compute H(f) and H(f 0 g). We have the commutative dia-
gram
f-l(a) x f-l(b)

gxg 1~
because
9 ¢(g(y)) - ¢(g(x)) ( ¢(y) - ¢(x) )
¢a,b(g(x),g(y)) = 11¢(g(y)) - ¢(g(x))11 = 9 1I¢(y) - ¢(x) II .

Consequently,
H(f 0 g) deg(g x g) = deg(g)H(f).
Here, we know that deg(g) = -1 (111.1.6(2), p. 100) and deg(g x g) = +1.
Thus, we conclude deg(g)H(f) = H(f 0 g).
Case IV: deg(g) = d > O. Replacing 9 with a homotopic mapping of the
type obtained in the proof of V.2.1, p. 191, we can assume that:

(1) The north pole p has a neighborhood V such that g-l(V) is a disjoint
union of d open sets Ui and on each of them 9 is a diffeomorphism
that preserves orientations.
(2) There are open sets Vi :J Ui , which do not meet the south pole, and
after projection from the south pole, every two of them can be sepa-
rated by a hyperplane.
(3) Off the Vi's everything goes to the south pole.

We denote by gi the mapping that coincides with 9 on Vi and maps ev-


erything else to the south pole. By construction, deg(gi) = +1. Now, let
a, b i= f( -p) be regular values of f. By a diffeotopy of §2m-l that col-
lapses everything to the north pole (11.6.3, p. 81), we can assume f-l(a)
and f-l(b) are contained in V, and so a, b are also regular values of fog,
both distinct from f(g( -p)) = f( -p). In this situation we have

(f 0 g)-l(a) = Ug;l(f-l(a)), (f 0 g)-l(b) = Ug;l(f-l(b)),


i i

so that
H(f 0 g) = £((f 0 g)-l(a), (f 0 g)-l(b))
= L£((fo9i)-1(a),(fogj)-1(b)).
i,j
202 v. The HopE Tlleorems

We look first at a pair i =I- j. Since the two inverse images involved are
projected into Ui and Uj, respectively, there is a hyperplane L that sepa-
rates those projections. Hence no line joining a point in one to a point in
the other can be parallel to L, and the corresponding map into §2m-2 is
not surjective and hence has degree o. Thus we are left with

H (f 0 g) = L e((f 0 gd -1 (a), (f 0 gi) -1 (b)) = L H (f 0 gd.


i i

As deg(gi) = +1, by Case II we know that H(f 0 gi) = H(f), and we


conclude that
H(f 0 g) = H(f) . d = H(f) deg(g).
Case V: deg(g) = d < O. This runs like Case IV, using Case III instead of
Case II, with the only modification that 9 reserves orientations on Ui and
consequently deg(gi) = -1. 0

By the preceding proposition, given f : §2m-1 -+ §m with H(f) =I- 0, its


compositions with mappings 9 : §2m-1 -+ §2m-1 of different degrees give
non-homotopic mappings fog: §2m-1 -+ §m. Thus we get an injection
Z = '7r2m-1 (§2m-1) -+ '7r2m-1 (§m).
Clearly, the best 1's for this construction seem to be those with H(f) =
±1. The classical method to define them follows.

(4.2) Hopf fibrations. Let F : ]Rm x ]Rm -+ ]Rm be the multiplication


of (i) complex numbers for m = 2, (ii) quaternions for m = 4, and (iii)
oct onions for m = 8, and define the Hopf fibration associated to F by

f : §2m-l -+ §m : (x, y) t--t (11x11 2 - IIYI12, 2F(x, y)).

Thus we obtain smooth mappings

whose Hopf invariant is 1. The fibers of these mappings are, respectively,


§l, §3, and §7; that is, the three mappings are fibrations of spheres by
spheres. We will content ourselves with the proof of this for m = 2, the
first case settled by Hopf in 1931 (see 1.3, p. 30).

Proposition 4.3. The Hopf invariant of the Hopf fibration §3 -+ §2 is 1.

Proof. The explicit equations of this mapping are


4. The Hop£ fibration 203

An easy computation shows that the poles of §2, a = (0, 0, -1) and b =
(0,0,1) (both =I- (-1,0,0) = j(O,O,O,-l)), are two regular values with
inverse images:

{
Ca = j-1(a) = §3 n {Y1 + X2 = Y2 + Xl = O},
Cb = j-1(b) = §3 n {Y1 - X2 = Y2 - Xl = O}.
By projection from the south pole we get in ]R3 two ellipses Ea and Eb:

E .
a'
{v(u-1)2+2v2=2,
+ w 0, = Eb .
.
{v(u+1)2+2v
- w 0, 2 =2.
=

They are depicted below, with the orientations (a and (b that they carry
as inverse images.
(3

Let us describe (a by means of II.7.6, p. 89. We localize j using (i) the


projection from the south pole in §3 and (ii) the projection from the north
pole in §2. Since both projections reverse orientation (II.7.5(3), p. 88), we
get the right orientation in the ellipse Ea, which for this localization has
the equations

91 : t(4u 2 + 4v 2 - ~w2 - t 2) = 0, t=1-U 2 - V 2 -W 2,


{ where {
92 - is (8uw - 4vt) - 0, 8 = (2 - t)2 - 8vw - 4ut.

The denominator 8 is important (at least its sign), because it affects the
derivative of the localization 9 = (91,92) of f. Let VI and V2 be the gra-
dients of 91 and 92, respectively. Then the inverse image is oriented by
204 V. The HopE Theorems

the vector product 111 x 112. Indeed, that vector product is perpendicu-
lar to the gradients, hence tangent to the inverse image, and of course
det(lII, 112, 111 X 112) > 0. In addition, dpg maps 111,112 onto a positive basis
in JR2 :

(Cauchy-Schwartz). With this settled, as orientation is determined by one


single point, we compute 111 and 112 for an easy one, namely (1 + y'2, 0, 0).
For this point, u > 0, v = W = 0, t = -2u < 0, 6> 0, and we have

{Ill:
112 -
t(8U(1- u)~,o) = (-,0,0),
;S(0,8u,8u) - (0,+,+),

so that the orientation (a is that of

as in the figure.
For (b the argument and computations are most similar, but in this case
we localize using in §2 the projection from the south pole, which preserves
orientation. Thus 111 x 112 does not give the right orientation, and we must
take the opposite -111 x 112. The details are left to the reader.
Finally, we have to compute the degree of the mapping

lC'2 ( ) q- p
~ : Ea x Eb -7 0) : p, q r-+ Ilq _ pil .

Now, the derivative of this mapping at (p, q) is quite simple: it is the linear
mapping (a, {3) r-+ {3 - a, followed by the orthogonal projection onto the
plane perpendicular to q - p, plus a scaling by the factor 1/ II q - pll. Then we
pick p,q and a,{3 as in the figure. It is clear that ~-l(~(p,q)) = {(p,q)};
hence we only must check that the two vectors

d(p,q)~(a, 0) = a', d(p,q)~(O, {3) = {3'

form a positive basis of Tcp(p,q)§2. This last linear space is the plane per-
pendicular to q - p, and a', {3' form a positive basis if and only if

det(q - p, a', {3') > 0.

But this should be clear from the picture! D


4. The Hopi fibration 205

As mentioned before, once we have this, composition with the Hopf


fibration gives an inclusion Z C 7r3(§2), which shows that 7r3(§2) is infinite.
As a matter of fact, it can be proved that the above inclusion is an equality;
hence
7r3(§2) = Z.

This is the first non-zero homotopy group of a sphere (excluding of course


7rm(§m) = Z), because 7rk(§I) = 0 for k > l.

Concerning the other two Hopf fibrations, it is also true that their Hopf
invariant is 1; hence we have injections 7r2m_I(§2m-l) --+ 7r2m_I(§m), and
so Z C 7r2m_I(§m), but no more: it is known that

Furthermore, except for m = 2, the Hopf invariant does not determine


the homotopy class: there are non-homotopic mappings §7 --+ §4 (and
§15 --+ §8) with the same Hopf invariant. On the other hand, not every
integer is a Hopf invariant for arbitrary even m (see 111.5.2(4)): even Hopf
invariants always occur, but odd ones occur only for the crucial dimensions
m = 2,4,8. All of this enters into the wide open problem of computing
homotopy groups of spheres, a beautiful topic which stands far beyond our
reach here.

Exercises and problems


Number 1. View the unit sphere §2m+1 in C m + 1 as defined by ZlZl + .. ·+Zm+lZm+l =
1. The mapping h: §2m+l -t ClP'm : Z = (Zl, ... ,Zl) -t (Zl : ... : Zm+l) is also called
Hopi fibration. Explain this name for m = 1. Use the fact that 7l"2m+l (§l) = 0 to show
that two continuous mappings I, g : §2m+1 -t §2m+ 1 such that hoi = hog must have the
same degree. Compute the degree of all a : §2m+1 -t §2m+l such that h(a(z)) = h(z).
Compare this with Problem Number 7 of 111.1.
Number 2. Let I: §2m-l -t §m and g : §m -t §m be smooth mappings (m ~ 2). Show
that
H(g 0 f) = deg(g/ H(f).
Number 3. Let I : §2m-l -t §m be a continuous mapping with non-zero Hopf invariant
(not necessarily 1). Let 9 : §2m-l -t §2m-l be a continuous lifting of g : §m -t §m (that
is, log = go f). Show that deg(g) = deg(g?, and deduce that few continuous mappings
are liftings. Deduce also that liftings of homotopic mappings are homotopic. What about
the other way around?
Number 4. Let g : §2m-l -t §2m-l and I: §2m-l -t §m be smooth mappings (m ~ 2).
Show that
fl(f 0 g) = fl(f) deg(g),
where fl is the invariant defined in Problem Number 1 of 111.5.
206 v. The HopE Theorems

Number 5. Let f : §2m-l --+ §m and 9 : §m --+ §m be smooth mappings (m ~ 2). Prove
the formula

Number 6. Compute the invariant (l(f) of the Hopf fibration f : §3 --+ §2.

5. Singularities of tangent vector fields


In this section we describe the information borne out by tangent vector
fields on a manifold. That information is disclosed by the computation of
the so-called index of a tangent vector field.
Let M c RP be an oriented, boundaryless, smooth manifold of dimen-
sion m.

(5.1) Non-degenerate zeros of a tangent vector field. A (smooth)


tangent vector field on M is a smooth mapping ~ : M -+ RP such that
~(z) E TzM for every z E M. We can represent ~ in local coordinates
as follows. Let 'P : U -+ M be a parametrization of M. For x E U set
z = 'P(x) E M. The tangent space TzM is generated by the basis of partial
derivatives
a
-a I a'P (x), 1 ~ i ~ m.
= dx'P(ei) = -a
Xi z Xi
Hence we can write

where the functions ~i : U -+ R are uniquely determined and smooth. We


will also consider the smooth mapping

~: U -+ R m : x H (6(x), ... ,~m(x)).

Now suppose z = 'P(x) is a zero (or singularity) of ~, that is, ~i(X) = 0 for
all i. Then

= ~ a~i (x) a'P (x) = ~ a~i (x)~1 .


~
i=l
ax'J ax'~ ~
i=l
ax'J ax'~ z
5. Singularities of tangent vector fields 207

Thus, we actually have a linear mapping dz~ : TzM --* TzM c '\R'P, whose
matrix with respect to the basis of partial derivatives is the Jacobian matrix
Jx1. of the localization f.. In particular, the determinant of dz~ is a well-
defined invariant of ~:

-= (a~i
Jx~ -a (x) ).. '
Xj ~,J

We will say that the zero z of ~ is non-degenerate when this determinant


is not zero. This is equivalent to saying that the localization 1. : U -+ lR.m
is a local diffeomorphism at x, hence locally injective, so that our zero z is
isolated. If z is non-degenerate, the sign ±1 of det(dz~) i= 0 is called the
index of ~ at z and is denoted by

Examples 5.2. Below there are some basic examples of vector fields on
the plane, with a single non-degenerate zero, and the corresponding index.
What is depicted is the flow generated by the field, which consists of the
integral lines whose tangents are prescribed by the field. They come from
an associated differential equation, and the investigation of their nature is
a highly interesting topic beyond our purposes here.

Source (or impulsor) Sink (or attractor)

~ = -xtx - yty ' ~ = (-x,-y),

- (-1 0)
d(o,o)~ = 0 -1 '

Saddle

~ = x tx - y t y ~ = (x, -y),

1 - (1
'

d(o.o)~ = 0 -1 '
0)
indo ~ = signo ~ = -1
208 V. The HopE Theorems

Circulation (or center)

But we want to deal with arbitrary isolated zeros, not only the non-de-
generate. This is achieved as follows.
Proposition and Definition 5.3 (Splitting of an isolated zero). Let ~ :
M -+ IRP be a smooth tangent vector field. Let il be an open set where ~
has a unique zero z. Then there are an open neighborhood U of z, with
U c il, and a smooth tangent vector field ( : M -+ IRP such that we have
the following:

(1) U is a smooth compact manifold with boundary Z = U \ U.


(2) (:= ~ off U.
(3) ( has finitely many zeros in il, all of them in U, and all of them
non-degenerate.

We say that ( is a splitting of ~ at z.

Proof. The question is clearly local; hence we can assume il = IRm , ~ = ~,


and z = O. Then, pick a regular value a of ~ with Iiall < 11~(x)11 for
1 ::; Ilxll ::; 2. Also, let () we a smooth bump function with 0 ::; () ::; 1, () == 1
on Ilxll ::; 1, and () == 0 on Ilxll ~ 2. Then set ( = ~ - a(). We have:

(i) (== ~ off U : Ilxll < 2 and hence has no zeros off U.
(ii) II~II > lIall ~ Ila(}11 on 1 ::; Ilxll ::; 2; hence ( has no zeros off V : IIxll <
1.

(iii) (= ~ -a on V; hence the zeros of (in V are the points in Vn~-l(a).


As a is a regular value of ~, those zeros are finitely many and non-
degenerate.

Finally, U is a closed ball in IRm , hence a compact manifold with boundary


the sphere Z : U \ U: Ilxll = 2. 0
5. Singularities of tangent vector fields 209

This splitting is the key step to defining the index at isolated degenerate
zeros:

Proposition and Definition 5.4. Let € : M -+ ffi.P be a smooth tangent


vector field, and let z E M be an isolated zero of €. Let ( be a splitting of
€ at z in the domain n corresponding to a parametrization tp. Then

where the sum ranges over all (non-degenerate) zeros z* of ( and S is a


small enough sphere centered at x = tp-l(z). Hence, this sum does not
depend on either the splitting or the parametrization.
This sum is the index of € at z. We write

If € has finitely many zeros, then all of them are isolated, and the sum
of the indices of € at the zeros is the total index of €. We write

(the sum is zero if the field has no zeros).

Proof. Consider the open neighborhood U c n associated to the splitting,


and its boundary Z = U \ U. Set X = tp-l(Z) and x* = tp-l(z*), so that

L ind

z• (() = L sign

x • (()
(by definition of index
at a non-degenerate zero)

= deg( 1I~lIlx) (by IV.4.5, p. 157)

= deg( II~II Ix) (because € == ( on Z)

II~IIIJ
(by the Boundary Theorem,
= deg( 111.3.3, p. 111)

Now, the right-hand side does not depend on (, and the left-hand side does
not depend on tp; hence none depends on either ( or tp. 0
210 V. The HopE Theorems

Examples 5.5. Here we give two examples of degenerate isolated zeros


with indices ±2:
(;=
~
-2xy JL
ax + (x 2 - y2)JL
ay'
Di ole
p
1 f = -L =
-

II~II
(
- 2xy, x
2
- y )
x 2 + y2
2

'
indo ~ = deg(flllxll=/O) = +2

Exercises and problems

Number 1. Show that the only possible indices at an isolated zero of a tangent vector
field on a curve are 0, +1, and -l.
Number 2. Produce smooth tangent vector fields on the plane jR2 with isolated zeros
of arbitrary indices.
Number 3. Let M be a compact, boundaryless, smooth manifold of dimension m.. Sup-
pose M is parallelizable, that is, there are tangent vector fields 6, ... , €m such that
{€I ,x, ... , €m,x} is a basis of Tx M at each point x E AI; this defines an orientation on PvI,
Show that:
(1) Each (smooth) tangent vector field € on M is defined by € = L::l J;€i, where
every Ii : AI -+ jR is a smooth function.
(2) A zero of € is a zero of the mapping 1= (/l, ... ,lm) : M -+ jRm, and it is
non-degenerate if and only if it is a regular point of I.
(3) If € has only non-degenerate zeros, then it has finitely many, and Ind(€) = deg(f).
Conclude that on a parallelizable compact manifold, all tangent fields with isolated
zeros have total index zero.
Number 4. Let € be a smooth tangent vector field on a smooth manifold AI, and let
I: M -+ jR be a smooth function. Consider the tangent vector field (. = I(z)€ •. Study
the zeros of ( in terms of those of €. When are they isolated? When non-degenerate?
Can something be said concerning indices in the latter case?
Number 5. Use the dipole (V.5.5 above) and stereographic projection to obtain a tan-
gent vector field on §2 with a unique zero of index +2.
Number 6. Consider the vector field on jR3 defined by
28 28 2 8
€ = xz 8x + yz 8y + z(z - 1) 8z'
6. Gradient vector fields 211

and let Melle be the surface x 2 +y2 = Z2 -1, Z > O. Show that erestricts to a tangent
vector field on M with a unique zero, and compute its index.
e
Number 7. Let (resp., () be a tangent vector field on a manifold M (resp., N) with
an isolated zero x E M (resp., YEN). Prove that ex ( is a tangent vector field on
M x N, that z = (x, y) is an isolated zero of ex (, and that

6. Gradient vector fields


We will discuss here a particular case of tangent vector fields and compute
their indices. We start with the following:

Proposition and Definition 6.1. Let Me lRP be a boundary less smooth


manifold of dimension m, and let f : M -+ lR be a smooth function. Then
there is a unique smooth tangent vector field ~ such that

dzf(u) = (u, ~z) for all u E TzM, Z E M,

where (., .) stands for the Euclidean scalar product in lRP .


This vector field ~ is called the gradient of f and is denoted grad(J).

Of course, if M is an open set of lRP , then grad(J) = (It, ... ,/!; ).


Proof. First recall that for every Z E M the mapping v f-t (., v) is a lin-
ear isomorphism from TzM onto its dual space C(TzM, lR) (this is the Riesz
Representation Theorem). Consequently, ~ is uniquely defined by the condi-
tion in the statement, and we must check that it is indeed a smooth tangent
vector field, which is a local matter. Thus, consider any parametrization 'P
of an open neighborhood U of z E M and the corresponding localization

and let us show that the ~k 's are smooth functions. We have
212 V. The HopE Theorems

Thus we obtain a linear system with unknowns the ~k 's, and the matrix
Gxtp of this system is the Euclidean Gramm matrix of the basis of partial
derivatives, which is the matrix of the Euclidean scalar product with respect
to that basis. Clearly such a Gramm matrix is positive definite and in
particular has determinant > O. Consequently, we can solve the system
and find smooth expresions for the ~k 'so We are done. 0

Note that the critical points of f : M -+ lR are exactly the zeros of its
gradient (and this is why zeros of tangent vector fields are called singular-
ities). We want to determine those which are non-degenerate zeros (V.5.1,
p. 206). Such a non-degenerate zero is called a non-degenerate critical point
of f.

(6.2) The Jacobian of a gradient vector field at a zero. Let f :


M -+ lR be a smooth function as above, and set ~ = grad(f). Let z be a
critical point of f, that is, a zero of ~. We consider a parametrization tp of
an open neighborhood U of z with, say, z = tp(x), and as in the preceding
proof we have

We are interested in the following Jacobian determinant:

Jx~- = (8~i
-8 (x) ) ...
Xj ~,J

Let us derive the formulas above:

8 2 (f 0 tp) = '""' 8(B£, l!;) ~k + '""' / 8tp, 8tp ) 8~k .


8x·8x·
J~
~ k 8x'J ~
k \8x·~ 8Xk 8x'J

Since z = tp(x) is a zero of ~, the first sum vanishes at x, and so


8 2 (f 0 tp) (x) = '""' / 8tp (x), 8tp (x)) 8~k (x).
8Xj 8Xi 7\ 8Xi 8Xk 8xj

Again we find the Gramm matrix Gxtp, and introducing the symmetric
matrix
Hx(f 0 tp) = (8 2 (f 0 tp) (x)) . . '
8xj8xi ~,J
we rewrite the sum in matricial form as follows:
6. Gradient vector fields 213

Since det( Gx<p) > 0, this completely determines the matrix Jx~, hence the
linear mapping dz~. Furthermore, we get
indz(~) = signdet(dz~) = signdet (Hx(f 0 <p)).
Here we include the case ind = sign = 0, meaning that z is a degenerate
zero of ~, that is, z is a degenerate critical point of f.

In the previous paragraph we encountered the Hessian of the localiza-


tion f 0 <po In fact, this is an intrinsic notion:
Proposition and Definition 6.3. In the setting above and with all of the
notation there, consider on TzM the quadratic form Qz(f) whose matrix
with respect to the basis {Ixt I
z} is Hx (f 0 <p). Let a tangent vector u E TzM
be defined by a curve germ ')'(t) in M, that is, ')'(0) = z, ')"(0) = u. Then
Qz(f)(u) = (f 0 ,),)"(0).
In particular, Qz(f) does not depend on the parametrization <po This
quadratic form is the Hessian of f at Z.

Proof. This is a straightforward computation. Let (U1,'" ,urn) be the co-


ordinates of u with respect to the basis of the partial derivatives. Then
2
Qz(f)(u) = 'L..J
" 0 (f 0 <p)
ox.ox. (x) Ui Uj.
ij t J

Now, set
= b1,' .. ,')'m),
1 = <p-1 0')'

so that the above coordinates are Ui = ')'~(O). We will compute (f 0,),)"(0).


First,

(f 0 ,),)'(t) = ((f 0 <p) 01)' (t) = L o(;;.<p) (1(t)h~(t),


. t
t

and then

(f 0 ,),)"(t) = '" 02(f 0 <p) (1(t)h'.(thHt) + '" o(f 0 <p) (1(t)h~'(t).


L..J
ij
axJ·ox·t J i L..J
t
ax,

For t = 0 we obtain
(f 0 ')'
)"()
0
",02(fo<p)()
= L..J ax .ox. x Uj Ui
ij J t

(the second sum vanishes because x = 1(0) is a zero of all the partial
derivatives 8U °If! ) ).
8Xi
0
214 V. The Hopi Theorems

The Hessian is used to define the index of a critical point:

Definition 6.4. Let f : M -+ lR be a smooth function and let z E M be


a non-degenerate critical point of f. The index (= the number of negative
eigenvalues) of the Hessian Q z (f) is called the index of f at z and is denoted
by indz(f).

With this we can compute indices of gradient vector fields:

Proposition 6.5. Let f : M -+ lR be a smooth function and let ~


grad (f) . Let z E M be a non-degenerate critical point of f. Then

Proof. Keeping all preceding notation, we already know that

indz(~) = signdet (Hx(f 0 cp)),


where Hx(f ocp) is the symmetric matrix of Qz(f) with respect to a suitable
base. Recall now that the determinant of a matrix is the product of its
eigenvalues. As indx (f) is precisely the number of those that are negative,
the assertion follows. D

(6.6) Morse functions. A Morse function f : M -+ lR is a smooth


function whose critical points are all non-degenerate, equivalently, such
that all the zeros of its gradient vector field ~ = grad(f) are non-degenerate.
Thus, they are isolated, and in case there are finitely many (for instance,
if M is compact), the total index of ~ is

where the z's are the critical points of f. This can be rewritten as

Ind(~) = ~) -l)kak'
k
where ak is the number of critical points with index k.
Of course, the question is whether or not Morse functions exist. The
answer is that they do. In fact, we will see that many a linear form is a
Morse function. We formulate this as follows.
6. Gradient vector fields 215

(6.7) Height functions. Let M c lRP be a boundaryless smooth manifold


of dimension m. For every a E lRP consider the smooth function

fa: M --+ lR: X t-+ (x,a).

This can be seen as a height function with respect to the linear hyperplane
H perpendicular to a: the oriented distance to H is given by

dist(x, H) = faIT fa(x).


Also, note that after a linear change of coordinates in lRP , we have a =
(0, ... ,0,1), and then fa(x) = xp.
Finally, denote by M the set of all a E lRP such that fa is a Morse
function.
The set M is quite explicitly determined for hypersurfaces:
Proposition 6.S. Let M c lRm+1 be a connected, closed, boundaryless,
smooth hypersurface, and let 1/ : M --+ §m be a unitary global normal field.
Let a E lRm+l. Then

fa : M --+ lR : X t-+ (X, a)

is a Morse function if and only if faITa E §m is a regular value of both 1/


and -1/.
In particular, the set M is a residual, hence dense, subset of lRm+l.

Proof. Recall that 1/ exists by III. 6.4, p. 129. On the other hand, since
f>..a = >'fa, we can assume II all = 1. Now, we look for the critical points z
of f = fa. Clearly, dzf = (., a)ITzM vanishes if and only if a is perpendicular
to TzM, that is, if and only if a = ±I/(z). Suppose this is true, and let us
see when the critical point z is non-degenerate. We first claim that

Indeed, let, be a curve germ in M with ,(0) = z, ,'(0) = u. Then we


derive
(J 0 ,)(t) = (,(t), a)
twice to get
Qz(J)(u) = (J 0 ,)"(0) = (,"(0), a).
But on the other hand, deriving (,'(t),I/(!(t») = 0, we see that
0= (,"(t) , I/(!(t») + (,'(t) , dzl/(!'(t») ,
216 v. The HopE Theorems

which for t = 0 gives

b"(O), v(z)) = (u, -dzv(u)).

But v(z) = ±a; hence b"(O), a) = (u, =r=dzv(u)) , and the claim follows.
Now, the claim implies that the quadratic form Qz(J) has, with respect
to any orthonormal basis of TzM, the same matrix as the self-adjoint en-
domorphism =r=dzv (III.2.6, p. 108), hence the same determinant. We have
thus seen that z is non-degenerate if and only if it is not a critical point of
±v. From this, the result follows readily.
The last assertion in the proposition is an easy consequence of the Sard-
Brown Theorem. D

For arbitrary manifolds one shows a slightly weaker result in a less


explicit way:

Proposition 6.9. Let M c 1l~.P be a boundaryless smooth manifold of di-


mension m. As above, let M be the set of all points a E ~p such that the
mapping
fa : M -t ~ : X H (X, a)
is a Morse function. Then M is dense in ~p.

Proof. First we remark that every point in M has a neighborhood U on


which some linear projection 1T : X H (Xii' ... ,Xim ) defines a local coordi-
nate system. Indeed, dp1T = 1T must be injective on TpM for some choice of
indices ik, and then the Inverse Function Theorem applies. Moreover, since
M has a countable basis of open sets, it has a countable covering consisting
of such open coordinate domains as U.
Now pick any U as above, and, to symplify notation, suppose the projec-
tion 1T corresponds to the indices ik = k; let X = (x', x") E ~p = ~m X ~p-m .
We claim that the following statement holds true:
Let a" E ~p-m, and denote by Mu(a") the set of all a' E ~m such
that
falu : U -t ~: x H (x', a') + (x", a")
is a Morse function. Then the set ~m \ Mu(a") has measure zero.
Assume this for the moment, and consider the set Mu of all a E ~p
such that falu is a Morse function. What the claim says is that ~p \
Mu cuts every (p - m)-plane x" = a" in a measure zero set; hence, by
the Fubini Theorem, ~p \ Mu itself has measure zero. Since there are
6. Gradient vector fields 217

countably many U's, the union Uu (~p \ Mu) also has measure zero, hence
has empty interior, and its complement nu Mu is dense. As, obviously,
M ::) nu Mu, we are done.
To complete the argument, let us prove the claim. Since 7rlu is a dif-
feomorphism onto an open set W C ~m, its inverse has the form x' H
(x', </>(x')), for a suitable smooth function </> : W -+ ~p-m. Then the local-
ization of lalu is

9a: W -+ ~: x' H (x', a') + h(x'), where h(x') = (</>(x'),a"),

and straight from the definition one sees that 9a is a Morse function if
and only if -a' is a regular value of the gradient grad(h) : W -+ ~m.
Consequently, the claim follows from the proof of the Sard-Brown Theorem
(see (*) on p. 64). 0

Exercises and problems


Number 1. Let M c RP be a smooth manifold and let F : RP -t R be a smooth
function; set f = FIM. Describe grad(f) in terms of grad(F) and orthogonal projections
into the tangent spaces of M.
Number 2. Is there any smooth function f : §2 -t R with gradient -y :x + x :y ?
Number 3. Let f (resp., g) be a Morse function on a manifold M (resp., N). Show
that
h : M x N -t R: (x, y) f-+ f(x) + g(y)
is a Morse function on M x N, and describe its critical points, as well as their indices,
in terms of those of f and g.
Number 4. Let f and g be Morse functions on two compact manifolds M and N,
respectively. Show that for any two positive real numbers a and b large enough, the
function
h : M x N -t R; (x, y) f-+ (a + f(x»(b + g(y»
is a Morse function on M x N. Describe the critical points of h and their indices.

Number 5. Let M = 80(3) C R 3X3 be the rotation group in R 3 , consisting of all 3 x 3


matrices x = (Xij) whose columns form a positive orthonormal basis of R3. This M is a
boundaryless, compact, smooth manifold of dimension 3. Show that

h(x) = 2xu + 3X22 + 4X33


is a Morse function on M and that it has four critical points with indices 0,1,2,3.

Number 6. Study when the function f ; RlP'7n -t R : (xo : ... ; x m ) f-+ 11;11 2 L:k CkX~ is
a Morse function. Then, describe its critical points and their indices.
218 V. The HopE Theorems

7. The Poincare-Hopf Index Theorem


Now that the notion of total index has been described, we can turn in this
last section to the main fact that the total index is an invariant of the
manifold, independent of the tangent vector field used to compute it. To
prove this, we must take a closer look at tubular neighborhoods.

(7.1) Normal vector fields on tubes. Let M c lRP be a compact


boundaryless smooth manifold of dimension m. Then, M has a tubular
retraction p : U ~ M (II.4.3, p. 71) such that x - p(x) is perpendicular to
Tp(x)M. Consider the smooth function
r: U ~ lR: X t--+ IIx - p(x)112.
As x - p(x) is perpendicular to dxp(lRP ) C Tp(x)M, it follows easily that
gradx(r) = 2(x - p(x)).
This gradient vanishes exactly for x = p(x), so that x E M and r(x) = O.
Thus, any c > 0 is a regular value of r, so that N : r(x) ~ c is a smooth
manifold with boundary aN : r(x) = c. We also know that gradx(r) is
perpendicular to aN at x, and consequently

) gradx(r) 1
TJ(x = II gradx(r)II = yIc(x - p(x))
is a unitary normal vector field on aN (III.2.6, p. 108).
Now recall that (i) Ilx - p(x) II = dist(x, M) for every x E U and (ii)
dist(M, lRP \ U) > O. Consequently, for c > 0 small enough
N = {x E lRP : dist(x, M) ~ vE}, aN = {x E lRP : dist(x, M) = vE}.
7. The Poincare-Hopi Index Theorem 219

Summing up, N is a compact manifold, and 'fJ is a normal vector field on


its boundary aN, which we call the normal vector field on a tube around
M.
Using the above normal vector fields, we can state the main theorem:
Theorem 7.2 (Poincare-Hopf Index Theorem). Let Me jRP be a compact
e
oriented boundaryless smooth manifold, and let be a smooth tangent vector
field with isolated zeros. Let'fJ: aN -+ §p- 1 be a normal vector field on a
tube around M. Then
Ind(e) = deg('fJ).

Proof. By V.5.3, p. 208, and V.5.4, p. 209, we can split all zeros of e to
assume they are non-degenerate. Once this is the case, with the notation
of the preceding paragraph, we define
f: N -+ jRP: X t-+ X - p(x) +e(p(x)).
We claim that 0 is a regular value of f. Indeed, since x - p(x) and e(p(x))
are perpendicular, f(x) = 0 if and only if x = p(x) and e(p(x)) = O. This
means that x E M is a zero a of e. Furthermore, given such a zero a, we
have:

(1) f == eon M; hence daflTa M = dae.


(2) p == a on the orthogonal complement a + E of TaM; hence daflE ==
IdE.

This means that daf = IdE El1dae, so that det(daf) = det(dae) =1= 0, since a
is a non-degenerate zero of e. This shows that 0 is a regular value of f, its
preimages are the zeros of e, and

a a

On the other hand, by IV.4.5, p. 157,

deg (1I~lIlaN) = Lsigna(J);


a

hence we only must show that deg('fJ) = deg (nfrrlaN)' But these two
mappings from aN into §p- 1 are homotopic because they do not have
antipodal images (V.1.4, p. 186). Indeed, if they had some, i.e., if
--X(x - p(x)) = x - p(x) + e(p(x)), with x E aN and -X > 0,
then e(p(x)) = -(1 + -X)(x - p(x)). This is impossible for x ~ M. 0
220 v. The HopE Theorems

Remark 7.3. If m is odd, this invariant has little significance. Indeed,


first note that Ind(-e) = (-l)mInd(e) (this is immediate for the index at
every non-degenerate zero and then follows for the total index). Thus, if
m is odd, Ind(e) = o. 0

(7.4) Euler characteristic. The Poincare-Hopf Theorem also says that


the total index is the Euler characteristic x(M) of M (hence = 0 if m is
odd) and consequently depends only on the topological type of M:

Ind(e) = X(M).
However, to bring in X, we would need a rigorous definition, which is not
our concern here. Instead we draw some pictures to see how X enters the
scene.
First notice that by the last theorem, it is enough to find a field e
whose total index is indeed the Euler characteristic. We will describe some
informal but very natural means to construct such a e.
(1) Suppose we have a triangulation. Then we mark one point ak at
each face of dimension k. These points will be the zeros of our field e, and
all are non-degenerate. In fact, e can be defined on each face by a source at
- k a
ak, that is, by the local form e = I:i=l Xi ax;. Of course, only on the faces
of maximal dimension k = m do we get a true source. On the others the
flow is incoming from higher-dimensional faces. Thus we start at the am '8
with sign +1, and the sign changes to -1 at the am-l's, and then again to
+1 at the am -2 's, and so on. Thus we have
Ind(e) = Am - A m- 1 + Am-2 - ... + (_l)m Ao
where Ak stands for the number of faces of dimension k. If m is even, this
sum is X, and if m is odd, it is -X, and we can change e to -e (although
it does not matter much because in that case X = 0).
7. The Poincare-HopE Index Theorem 221

The preceding picture shows X in dimension 2: we get a source at every


face, a saddle at every edge, and a sink at every vertex, adding up to the
classical Euler formula X = F - E + V.
(2) We can imagine the construction above in a physical way, as the
flow of a liquid poured over the manifold from some vintage points. We
can use this image to guess the Euler characteristic of a torus with 9 holes.
The figure below shows this: sources have index +1, saddles -1, and sinks
+1.
t t
source: +1

sink: +1

E=O E=-2 E=O E=-2

When we add up all indices, the four zeros around each hole cancel each
other, and each pair of zeros in between two consecutive holes gives -2.
Since there are 9 holes, there are 9 - 1 of those pairs, and the final result
is what it must be:

X = Ind = Lind = - 2(g - 1) = 2 - 2g.

This 9 is the genus of the surface. The reader can try different forms of
pouring the liquid and check that, of course, the result is always the same.
(3) For the sphere §2 C IR3 (g = 0), we have just a source and a sink,
which gives index 2 and X(§2) = 2 = 2 - 2g. Or we can simply use the field
e(x, y, z) = (-y, x, 0) depicted after the Hedge-hog Theorem, III.7.4, p. 135:
its two zeros are circulations and hence have index +1. Also, we understand
now what that theorem says for arbitrary compact, boundaryless manifolds:
if there is a tangent vector field e without zeros, then X = Ind(e) = o.

(7.5) The Morse inequalities. This is a description of the Euler charac-


teristic by means of gradient vector fields. Again let M be an oriented, com-
222 V. The HopE Theorems

pact, boundaryless, smooth manifold of dimension m, and let f : M --+ lR


be a Morse function (which always exists; V.6.S, p. 215, and V.6.9, p. 216).
We have already computed the total index of the gradient ~ = grad(f) of
f:
Ind(~) = 2)
-llok,
k

where Ok is the number of critical points of index k of f (V.6.6, p. 214).


Thus we get
X(M) = ~)-l)kok'
k
This formula for the Euler characteristic in terms of the critical points of a
Morse function is part of the so-called Morse inequalities (in fact this is the
unique equality among them). Those inequalities are actually formulated
in terms of the homology groups of M and escape our context.

(7.6) The Gauss-Bonnet Formula. As we explained earlier (111.2.6, p.


lOS), this is the computation of the integral curvature

for a compact hypersurface M C lRm +1 of even dimension m with Gauss


mapping v. There we saw how the first equality comes from degree theory
by integration, and what was left was the computation

deg(v) = !X(M).
But this follows from the Poincart3-Hopf Index Theorem, V.7.2, p.219,
which for hypersurfaces of even dimension has a better form. Let us sketch
the argument.
First notice that as M is a hypersurface of dimension m, aN consists
of two disjoint copies M+ and M_ of M: for every x E M there are exactly
two points x+ and x_ at a distance V€ from x, and the signs can be chosen
by the conditions

x+ = x + V€v(x), hence 1J(x+) = v(x),


{
x_ = x - V€v(x), hence 1J(x_) = -v(x).
Thus p induces two diffeomorphisms M+, M_ --+ M with inverses described
just above. In particular, the two derivatives of those inverses are

arbitrarily close to Id for E small enough.


7. The Poincare-HopI Index Theorem 223

All of this readily implies that p preserves orientations on M+, and it does
not on M_. Finally, write

where 0- is the antipodal diffeomorphism of §m. Since the diffeomorphism


0- has degree +1 if m is odd and -1 if m is even (III.1.6(1), p. 99), we get

deg('T]) =deg('T]IM+) + deg('T]IJ\,L)


= deg(v 0 PIM+) + deg(o- 0 v 0 pIM_)

= deg(v)( +1) + (±1) deg(v)( -1) = {o


2deg(v)
if m is odd,
if m is even.

As we have already explained, deg('T]) = X(M), and for m odd we again


get that X(M) = 0 and we get nothing about deg(v). But for m even we
obtain X(M) = 2 deg(v), and the Gauss-Bonnet Formula follows.

Exercises and problems


Number 1. Let §m C ]Rm+l be a sphere of even dimension. Use the tangent vector field

to compute the Euler characteristic X(§m).

Number 2. Prove the following product formula for the Euler characterisitic

X(M x N) = X(M) . X(N).

Number 3. Show that the torus T C ]R3 generated by the circle y = 0, (x - 2? +Z2 = 1
around the axis x = y = 0 is diffeomorphic to the product §l x §l C ]R4, and deduce that
224 v. The HopE Theorems

its Euler characteristic is O. Confirm this in three other ways:


(1) Obtain a tangent vector field on T without zeros, and apply the Poincare-Hopf
Index Theorem.
(2) Compute the degree of the Gauss mapping 1/ : T -+ §2, and apply the Gauss-
Bonnet Theorem.
(3) Exhibit a Morse function on T, and use the Morse Inequalities.
Notice that (1) and (3) can also be done for §1 x §1 C lR\ but (2) cannot, as the
latter is not a hypersurface.
Number 4. Prove, using Euler characteristics, that a sphere of even dimension cannot
be homeomorphic to a product of two other spheres.
Number 5. Let M C lR 3 be a compact smooth surface, and let f : M -+ lR be a Morse
function. Show that f has at least two critical points, and if there are no more, then M
has genus 0 (hence M is diffeomorphic to the sphere §2).
Number 6. Compute the Euler characteristic of the rotation group M = 80(3) C lR3X3
using the Morse function h in Problem Number 5 of V.6.
Number 7. Let M C lR 3 be the cornered sphere X4 + y4 + Z4 = 1. Check the Gauss-
Bonnet Theorem (i) through explicit computation of the integral curvature and (ii) com-
puting the degree of the Gauss mapping.
Names of mathematicians cited
Aleksandrov, Pavel Sergeevich Jacobi, Carl Gustav Jacob
(1896-1982), 32 (1804-1851),7
Amann, Herbert, 39 Jodel, Jerzy, 46
Jordan, Camille (1838-1922), 15
Bernstein, Sergei Natanovich
(1880-1968),34 Kronecker, Leopold (1823-1891), 7
Betti, Enrico (1823-1892), 13
Bohl, Piers (1865-1921), 13 Leray, Jean (1906-1998), 33
Brouwer, Luitzen Egbertus Jan Liouville, Joseph (1809-1882), 2
(1881-1996),14
Brown, Arthur Barton, 35 Marzantowicz, Waclaw, 46
Massabo, Ivar, 43
Caccioppoli, Renato (1904-1959), 34 Miranda, Carlo, 28
Cauchy, Augustin Louis
(1789-1857),2 Nagumo, Mitio, 35
Cech, Eduard (1893-1960), 30 Nirenberg, Louis, 43

Deimling, Klaus, 40 Ostrowski, Alexander (1893-1986), 2


Dyck, Walter Franz Anton
Picard, Charles Emile (1856-1941),
(1856-1934), 31
12
Dylawerski, Grzegorz, 46
Poincare, Jules Henri (1854-1912),
Elworthy, K. David, 43 11
Pontryagin, Lev Semenovich, 43
Fuhrer, Lutz, 38
Fuller, F. Brock, 45 de Rham, Georges (1903-1990),37
Riemann, Friedrich Bernhard Georg
Gauss, Karl-Friedrich (1777-1855), 2 (1826-1866), 13
Geba, Kazimierz, 43 Romero Ruiz del Portal, Francisco,
45
Hadamard, Jacques Salomon Rothe, E., 29
(1865-1963), 14 Rybicki, Slawomir, 48
Heinz, Erhard, 38
Hermite, Charles (1822-1901), 7 Sard, Arthur (1909-1980), 35
Hopf, Heinz (1894-1971), 28 Schauder, Juliusz Pawel
Hurewicz, Witold (1904-1956), 30 (1899-1943),33
Schmidt, Erhard (1876-1959), 28
Ize, Jorge, 43
-
225
226 Names

Schonfiies, Arthur Moritz


(1853-1928),15
Schwartz, Laurent (1915-2002), 38
Smale, Stephen, 42
Sturm, Jacques Charles Franr;ois
(1803-1855),2
Sylvester, James Joseph
(1814-1897), 7

Thorn, Rene (1923-2002), 42


Tromba, Anthony J., 43

Veblen, Oswald (1880-1960), 15


Vignoli, Alfonso, 43

Weierstrass, Karl Theodor Wilhelm


(1815-1897), 10
Weiss, Stanley A., 39
Historical references
We list below the references mentioned in Chapter I. We also include three
basic papers on the history of degree theory: [Siegberg 1980a], [Siegberg
1980bJ, and [Mawhin 1999J.

[Adams 1969] J.F. Adams: Lectures on Lie Groups. New York-Amsterdam: W.A.
Benjamin, Inc., 1969.
[Aleksandrov-Hopf 1935] P. Aleksandrov, H. Hopf: Topologie. Berlin: Verlag von
Julius Springer, 1935.
[Amann-Weiss 1973] H. Amann, S.A. Weiss: On the uniqueness of topological
degree. Math. Z. 130 (1973), 39-54.
[Banach 1922] S. Banach: Sur les operations dans les ensembles abstraites et leurs
applications aux equations integrales. Fundamenta Math. 3 (1922), 133-
181.
[Bohl1904] P. Bohl: Uber die Bewegung eines mechanischen Systems in der Nahe
einer Gleichgerwichtslage. J. Reine Angew. Math., 127 (1904), 179-276.
[Brouwer 1912a] L.E.J. Brouwer: Uber abbildung von Mannigfaltigkeiten. Mat.
Ann. 71 (1912), 97-115.
[Brouwer 1912b] L.E.J. Brouwer: On looping coefficients. Konink. Nederl. Akad.
von Wet. Proc. 15 (1912), 113-122.
[Brouwer 1912c] L.E.J. Brouwer: Continuous one-one transformations of surfaces
in themselves. CNAG Proc. 15 (1912), 352-360.
[Brouwer 1976] L.E.J. Brouwer: Collected works, vol. II. Geometry, analysis,
topology and mechanics. H. Freudenthal (ed.). Amsterdam-Oxford: North
Holland, 1976.
[Brown 1935] A. B. Brown: Functional dependence. Trans. Amer. Math. Soc. 38
(1935), 379-394.
[Caccioppoli 1936] R. Caccioppoli: Sulle corrispondenze funzionalli inverse dira-
mata: teoria generale e applicazioni ad alcune equazioni funzionali non lin-
eari e al problema di Plateau, I & II. Rend. Accad. Naz. Linzei 24 (1936),
258-263 & 416-421.
-227
228 Historical references

[Cauchy 1837aJ A.-L. Cauchy: Calcul des indices des fonctions. J. Ecole
Poly technique XV (1837), 176-226.
[Cauchy 1837bJ A.-L. Cauchy: Extrait d'une lettre sur une memoire publie it
Thrin, Ie 16 Juin 1833, et relatif aux racines des equations simultanees.
Comptes Rendus Acad. Bc. Paris IV (1837), 672-675.
[Cauchy 1855J A.-L. Cauchy: Sur les compteurs logarithmiques. Comptes Rendus
Acad. Bc. Paris XL (1855), 1009-1016.
[Cech 1932J E. Cech: Hoherdimensionale Homotopiegruppen. Verhandlungen der
Internationalen Mathematiken Kongresses, Zurich (1932), 203.
[Dancer 1983J E.N. Dancer: On the existence of zeros of perturbed operators.
Nonlinear Anal. 7 (1983), 717-727.
[Dancer 1985J E.N. Dancer: A new degree for §l-invariant gradient mappings and
applications. Ann. Inst. Henri Poincare, Anal. Non-lineaire 2 (1985), 329-
370.
[Dancer et al. 2005J E.N. Dancer, K. Geba, S.M. Rybicki: Classification of homo-
topy classes of equivariant gradient maps. Fundamenta Math. 185 (2005),
1-18.
[Deimling 1985J K. Deimling: Non-linear functional analysis. Berlin: Springer-
Verlag, 1985.
[Dyck 1888J W.F.A. Dyck: Beitrage zur Analysis situs I. Aufsatz ein- und zwei-
dimensionale Mannigfaltigkeiten. Math. Ann. 32 (1888), 457-512.
[Dyck 1890J W.F.A. Dyck: Beitrage zur Analysis situs II. Aufsatz Mannig-
faltigkeiten von n-dimensionen. Math. Ann. 34 (1888), 273-316.
[Dylawerski et al. 1991J G. Dylawerski, K. Geba, J. Jodel, W. Marzantowicz: An
§l-equivariant degree and the Fuller index. Ann. Pol. Math. 52 (3) (1991),
243-280.
[Elworthy-Tromba 1970aJ K.D. Elworthy, A.J. Tromba: Differential structures
and Fredholm maps on Banach manifolds. In Global Analysis (Berkeley,
1968) 45-94, Proc. Symp. Pure Math. 15. Providence, R.I.: American
Mathematical Society, 1970.
[Elworthy-Tromba 1970bJ K.D. Elworthy, A.J. Tromba: Degree theory on Banach
manifolds. In Nonlinear Functional Analysis (Chicago, 1968) 75-94, Proc.
Symp. Pure Math. 18, part 1. Providence, R.I.: American Mathematical
Society, 1970.
[Fuhrer 1971J L. Fuhrer: Theorie des Abbildungsgrades in endlich-dimensionalen
Raumen. Dissertation, Freie Universitat Berlin, 1971.
[Fuhrer 1972J L. Fuhrer: Ein elementarer analytischer Beweis zur Eindeutigkeit
des Abbildungsgrades in jRn. Math. Nachr. 54 (1972), 259-267.
[Fuller 1967J B. Fuller: An index of fixed point type for periodic orbits. American
J. Math. 89 (1967), 133-148.
Historical references 229

[Gauss 1813] K.-F. Gauss: Theoria attractionis corporum sphaeroidicorum el-


lipticorum homogeneorum methodo nova tractata. Gottingische gelehrt
Anzeigen, April 5 (1813), 545-552.
[Gauss 1833] K.-F. Gauss: Zur mathematischen Theorie der electrodynamis-
che Wirkungen. In Werke, 5. G6ttingen: K6niglichen Gesellschaft der
Wissenchaften zu G6ttingen (1867),601-626.
[Geba et al. 1986] K. Geba, I. MassabO, A. Vignoli: Generalized topological de-
gree and bifurcation. In Non-linear Functional Analysis and Applications
(Proc. NATO Adv. Study Inst., Maratea, Italy, 1985) 55-73, Mathematical
and Physical Sciences 173. Dordrecht: Reidel, 1986.
[Geba et al. 1990] K. Geba, I. MassabO, A. Vignoli: On the Euler characteristic
of equivariant gradient vector fields. Boll. Un. Mat. Ital. A(7) 4 (1990),
243-251.
[Hadamard 1910] J.-S. Hadamard: Sur quelques applications de l'indice de Kro-
necker. In [Tannery 1910], 437-477.
[Heinz 1959] E. Heinz: An elementary analytic theory of the degree of mapping
in n-dimensional space. J. Math. and Mech. 8 (1959), 231-247.
[Hopf 1925] H. Hopf: Uber die Curvatura integra geschlossener Hyperfliichen.
Math. Ann. 95 (1925), 340-367.
[Hopf 1926a] H. Hopf: Abbildungsklassen n-dimensionaler Mannigfaltigkeiten.
Math. Ann. 96 (1926), 209-224.
[Hopf 1926b] H. Hopf: Vectorfelder in n-dimensionalen Mannigfaltigkeiten. Math.
Ann. 96 (1926), 225-250.
[Hopf 1931] H. Hopf: Uber die Abbildungen der dreidimensionalen Sphiire auf die
Kugelfliiche. Math. Ann. 104 (1931), 637-665.
[Hopf 1933] H. Hopf: Die Klassen der Abbildungen der n-dimensionalen Polyeder
auf die n-dimensionale Sphiire. Comment. Math. Helvetici 5 (1933),39-54.
[Hopf 1935] H. Hopf: Uber die Abbildungen von Sphiiren auf Sphiiren niedrigerer
Dimension. Fundamenta Math. 25 (1935),427-440.
[Hopf 1966] H. Hopf: Ein Abschnitt aus der Entwicklung der Topologie. Jber.
Deutsch. Math.- Verein. 68 (1966), 96-106.
[Hurewicz 1935a] W. Hurewicz: Beitriige zur Topologie der Deformationen I.
H6herdimensionalen Homotopiegruppen. Proc. Akad. Wetenschappen 38
(1935), 112-119.
[Hurewicz 1935b] W. Hurewicz: Beitriige zur Topologie der Deformationen II.
Homotopie und Homologiegruppen. Proc. Akad. Wetenschappen 38 (1935),
521-528.
[Hurewicz 1936a] W. Hurewicz: Beitriige zur Topologie der Deformationen III.
Klassen und Homologietypen von Abbildungen. Proc. Akad. Wetenschap-
pen 39 (1936), 117-126.
230 Historical references

[Hurewicz 1936b] W. Hurewicz: Beitrage zur Topologie der Deformationen IV.


Aspharische Raume. Proc. Akad. Wetenschappen 39 (1936), 215-224.
[Ize 1981] J. Ize: Introduction to bifurcation theory. In Differential Equations
(Sao Paulo, 1981), 145-203. Lecture Notes in Math. 957. Berlin-New York:
Springer-Verlag, 1982.
[Ize et al. 1986] J. Ize, I. Massabo, A. Vignoli: Global results on continuation and
bifurcation for equivariant maps. In Non-linear Functional Analysis and
Applications (Maratea, Italy, 1985), 75-111. NATO Adv. Sci. Inst. Ser. C
Math. Phys. Sci. 173. Dordrecht: Reidel, 1986.
[Ize et al. 1989] J. Ize, I. Massabo, A. Vignoli: Degree theory for equivariant maps,
I. Trans. Amer. Math. Soc. 315 (2) (1989), 433-510.
[Ize et al. 1992] J. Ize, I. Massabo, A. Vignoli: Degree theory for equivariant maps,
II: The general §l-actions. Memoirs Amer. Math. Soc. 481 (1992).
[Iz€-Vignoli 2003] J. Ize, A. Vignoli: Equivariant Degree Theory. de Gruyter Series
in Nonlinear Analysis and Applications 8. Berlin: Walter de Gruyter & Co.,
2003.
[Jordan 1893] C. Jordan: Cours d'analyse de l'Ecole Poly technique, f. 1893.
[Kronecker 1869a] L. Kronecker: Uber Systeme von Functionen mehrerer Vari-
abeln, I. Monatsberichte koniglich Preuss. Akad. Wissens. Berlin, March 4
(1869), 159-193.
[Kronecker 1869b] L. Kronecker: Uber Systeme von Functionen mehrerer Vari-
abeln, II. Monatsberichte koniglich Preuss. Akad. Wissens. Berlin, August
5 (1869), 688-698.
[Leray-Schauder 1933] J. Leray, J.P. Schauder: Topologie et equations fonction-
nelles. Comptes Rendus Acad. Sc. Paris 1197 (1933), 115-117.
[Leray-Schauder 1934] J. Leray, J.P. Schauder: Topologie et equations fonction-
nelles. Ann. Sc. Ecole Normale Sup. Paris 51 (1934),45-78.
[Liouville-Sturm 1837] J. Liouville, J.-Ch.-F. Sturm: Note sur un tMoreme de
M. Cauchy relatif aux racines des equations simultanees. Comptes rendus
Acad. Sc. Paris 40 (1937), 720-724.
[Mawhin 1999] J. Mawhin: Leray-Schauder degree: A half century of extensions
and applications. Topol. Methods Nonlinear Anal. 14 (1999), 195-228.
[Miranda 1940] C. Miranda: Un'osservazione su un teorema di Brouwer. Boll. Un.
Mat. ftal. (2) 3 (1940), 5-7.
[Nagumo 1951a] M. Nagumo: A theory of degree of mapping based on infinitesi-
mal analysis. American J. Math. 73 (1951),485-496.
[Nagumo 1951b] M. Nagumo: Degree of mapping in convex linear topological
spaces. American J. Math. 73 (1951), 497-511.
Historical references 231

[Nirenberg 1971] L. Nirenberg: An application of generalized degree to a class


of nonlinear problems. In T'roisieme Colloque sur l'Analyse Fonctionnelle
(Liege, 1970), 57-74. Louvain: Vander, 1971.
[Picard 1891] Ch.-E. Picard: Sur Ie nombre des racines communes it plusieurs
equations simultanees. Comptes Rendus Acad. Sc. Paris 113 (1891), 256-
358.
[Picard 1891/1905] Ch.-E. Picard: TraiU d'Analyse, I and II. Paris: Gauthier-
Villars, 1891 and 1893/1905.
[Poincare 1883] J.H. Poincare: Sur certains solutions particulieres du probleme
des trois corps. Comptes Rendus Acad. Sc. Paris 97 (1883), 251-252.
[Poincare 1885a] J .H. Poincare: Sur les courbes definies par les equations dif-
ferentielles. J. Math. Pures Appl. 1 (1885), 167-244.
[Poincare 1895b] J.H. Poincare: Analysis Situs. J. Ecole Poly technique (2) 1
(1895), 1-123.
[Poincare 1886] J.H. Poincare: Sur les courbes definies par une equation dif-
ferentielle. J. Math. Pures Appl. 2 (1886), 151-217.
[Poincare 1899] J.H. Poincare: Complement it l'Analysis Situs. Rend. Circ. Mat.
Palermo 13 (1899), 285-343.
[Poincare 1900] J.H. Poincare: Complement it l'Analysis Situs. Proc. London
Math. Soc. 32 (1900), 277-308.
[Pontryagin 1955] L.S. Pontryagin: Smooth manifolds and their application in
homotopy theory. Amer. Math. Soc. Transl. (2) 11 (1959), 1-114, from
Trudy Mat. Inst. im Steklov 45 (1955).
[de Rham 1955] G. de Rham: VarieUs difJerentiables. Paris: Hermann, 1955.
[Rothe 1936] E. Rothe: Uber Abbildungsklassen von Kugeln des Hilbertschen
Raumes. Compos. Math. 4 (1936), 294-307.
[Ruiz del Portal 1991] F.R. Ruiz del Portal: Teoria del grado topol6gico gene-
ralizado y aplicaciones. Dissertation. Madrid: Universidad Complutense,
1991.
[Ruiz del Portal 1992] F.R. Ruiz del Portal: On the additivity property of the
generalized degree. Math. Japonica 37 (1992), 657-664.
[Rybicki 1994] S. Rybicki: A degree for §l-equivariant orthogonal maps and its
applications to bifurcation theory. Nonlinear Anal. 23 (1994),83-102.
[Sard 1942] A. Sard: The measure of critical points of differentiable maps. Bull.
Amer. Math. Soc. 48 (1942), 883-897.
[Schonflies 1902] A.M. Schonflies: Uber einen grundlegenden Satz der Analysis
Situs. Gott. Nachr. Math. Phys. Kl. (1902), 185-192.
[Siegberg 1980a] H. W. Siegberg: Brouwer degree: History and numerical com-
putation. In Numerical solutions of highly nonlinear problems, W. Forster
(ed.), 389-411. Amsterdam: North-Holland, 1980.
232 Historical references

[Siegberg 1980bJ H. W. Siegberg: Some historical remarks concerning degree the-


ory. American Math. Monthly 88 (1981), 125-139.
[Smale 1965J S. Smale: An infinite dimensional version of Sard's Theorem. Amer-
ican J. Math. 87 (1965), 861-866.
[Sylvester 1853J J.J. Sylvester: On a theory of the syzygetic relations of two ratio-
nal integral functions, comprising an application to the theory of Sturm's
functions, and that of greatest algebraic common measure. Philos. Trans.
Roy. Soc. London 143 (1853), 407-562.
[Tannery 1910J J. Tannery: Introduction Ii la theorie des fonctions d'une variable,
t. 2, 2ieme edition. Paris: Hermann, 1910.
[Thom 1954J R. Thom: Quelques proprif~tes globales des varietes differentiables.
Comment. Math. Helvetici 28 (1954), 17-86.
[Veblen 1905J O. Veblen: Theory of plane curves in non-metrical analysis situs.
Trans. Amer. Math. Soc. 6 (1905), 83-98.
Bibliography
For prerequisites, we recommend the quite ad hoc texts [4], [10], [16], and
[18]. On the other hand, there is a wealth of literature on degree theory
and related topics; we suggest as further reading the following books:

[1] J. Cronin: Fixed points and topological degree in nonlinear analysis.


Mathematical Surveys 11. Providence, R.I.: American Mathematical
Society, 1964.

[2] K. Deimling: Nonlinear Functional Analysis. Berlin: Springer-Verlag,


1985.

[3] A. Dold: Teoria de punto fijo (I, II, III). Mexico: Monografias del
Instituto de Matematica, 1986.

[4] J.M. Gamboa, J.M. Ruiz: Iniciacion al estudio de las variedades


diferenciales (2a edicion revisada). Madrid: Sanz y Torres, 2006.

[5] A. Granas, J. Dugunji: Fixed Point Theory. Springer Monographs in


Mathematics. New York: Springer-Verlag, 2003.

[6] V. Guillemin, A. Pollack: Differential Topology. Englewood Cliffs,


N.J.: Prentice Hall, Inc.,1974.

[7] M.W. Hirsch: Differential Topology. Graduate Texts in Mathematics


33. Springer-Verlag, New York-Heidelberg: Springer-Verlag, 1976.

[8] M.A. Krasnosel'skii: Topological methods in the theory of nonlinear


integral equations (translated from Russian). New York: Pergamon
Press, 1964.

[9] W. Krawcewicz, J. Wu: Theory of degrees with applications to bifur-


cations and differential equations. New York: John Wiley & Sons,
1997.

[10] S. Lang: Differential manifolds. Berlin: Springer-Verlag, 1988.


-
233
234 Bibliography

[11] E.L. Lima: IntrodUf;ao Ii Topologia Diferencial. Rio de Janeiro: In-


stituto de Matematica Pura e Aplicada, 1961.

[12] N.G. Lloyd: Degree Theory. Cambridge Tracts in Mathematics 73.


Cambridge: Cambridge University Press, 1978.

[13] 1. Madsen, J. Tornehave: From calculus to cohomology: de Rham


cohomology and characteristic classes. Cambridge, New York, Mel-
bourne: Cambridge University Press, 1997.

[14] J. Milnor: Topology from the differentiable viewpoint. Princeton Land-


marks in Mathematics. Princeton, N.J.: Princeton University Press,
1997.

[15] L. Nirenberg: Topics in nonlinear functional analysis (with a chapter


by E. Zehnder and notes by R.A. Artino). Courant Lecture Notes
in Mathematics 6. New York: Courant Institute of Mathematical
Sciences, American Mathematical Society, 1974.

[16] E. Outerelo, J .M. Ruiz: Topologia Diferencial. Madrid: Addison-


Wesley, 1998.

[17] P.H. Rabinowitz: Theorie du Degre Topologique et applications Ii des


problemes aux limites non lineaires (redige par H. Berestycki). Lec-
ture Notes Analyse Numerique Fonctionelle. Paris: Universite Paris
VI, 1975.

[18] M. Spivak: Calculus on manifolds: A modern approach to classical


theorems of advanced calculus. Boulder: Westview Press, 1971.
Symbols
y = f(x) 1 9 = 'ljJ 0 go cp-l : IKI --7 ILl 24
zn + alZ n - l + ... + an = ° 2 d(j) 25
J:~(j) 3 F(t, x) = IM~~l=g=g~1I 27
Z(Z) = X(X, y) + iY(x, y) 3 §+ 28
1f
211"i r
Z'{z)d
Z{z) z 3 f(Kl' K 2 ) 28
N = XI
Xo
JYl
Yo
(.d) 3 X(M) 30
~8 = /-Ll -/-L2 4 <Ph(X) = x - Fh(X) 34
aF aF d(<p, W,O) = d(<Ph' WM,O) 34
ax ay
w= 5 x-F(x)=O 34
£l. £l.
ax ay dist 36
w(r, a) = 2;i fr z~a 5
d(j,G,a) 36
w(j(r), O) = 2;i ff{r) ~ 6 fIRm <p(lxl)dx =1 38
w(j(r), O) = Lk w(r, ak)ak 6 d(y(x), n, z) 38
w(j(r) , 0) = Lkak 6 fn<p(ly(x) - zI)J(y(x))dx 38
E(k, f) 8 M(W) 41
A(k, f) 8 Mo(n) 41
#E(k, f) - #A(k, f) 8 an= n\n 41
X(Fo, F l ,· .. ,Fn) 8 f, foe, f* 43
H#E(k,f) - #A(k,f)) 8 d*(j, U,O) 44
V() X-XC)
11 d*(j, U) 44
Z = Ilx-xoll 3
11 p : §l --7 GL(V) 46
fFo=O(V,v)dS
w=Nl -N2 17 21, a = (ark::o 46

75" = {x E]Rn : Ilxll ~ 1} 20 Deg(j, n) 46


a75" = §n-l 20 Deg(j, n) = E(d*(j, n)) 47
• §l *a 47
K,IKI,IKI 23

-
235
236 Symbols

Deg(f,il) = (a r ) 47 {(x, u) E 11M: Ilull < r(x)} 72


8kf 49 H(t, x), Ht(x), Ht 75
8X'l··· 8x'k
54 [M,N] 75
'P = ('Pl, ... , 'Pp)
7rk(N) = [§k, N] 75
x = (Xl, ... ,xm) 54
'l/J-l 0 'P 7rk(M) = [M, §k] 75
54
dimx(M), dim(M) 54 Ilf(x) - g(x)11 < c(x) 76
(m 85
Un M = f- 1(0) 55
(M = {(x: x E M} 85
(8 fi ) 55
8xj
[8~1 Ix'···' 8~rn IJ = (x 85
h(U n M) = V n (l~m X {O}) 55
(MxN = ((M,(N) 86
lHIm : A ~ 0 56
signx(f) 86
8M 56
II : M --7 JR. m+1 87
8([0,1] x M) = Mo U Ml 57
8( 91
Tx M 58
EXEf-l(a) signx(f) 95
..iLl
8x, x -
- ~(
8x, x ) 58
deg(f) 95
gradx(f) 58
deg(h x ... x fr) 102
dxf(u) = (gradx(f), u) 58
T m = §1 x ... X §1 102
T(x,y)(M x N) = TxM x TyN 58
7r : §m --7 :mpm 103
dxf : TxM --7 TyN 59
r~(M) 104
dag = (d b'l/J)-1 0 dxf 0 da'P 59
d : r~(M) --7 r~+I(M) 104
1" (0) = do l' (1 ) 60
d(a 1\,8) 104
dxf(u) = (f 0 1')'(0) 60
dw = 0, w = da 104
JR.lP'm,Cr 61
H~(M,JR.) 104
TM 61
H~(M,Z) 104
VM 61
H;'(JR.m,JR.) = JR. 104
Cf , Rf 62
Hm(JR.m,JR.) = 0 104
Rf=N\f(Cf ) 62
fMda = 0 104
Txf-l(a) = ker(dxf) 62
fM : H;'(M, JR.) --7 JR. 104
O(n) 66
h = E"!'
.=1 !!..&
8x, 105
OM 66
w t-+ f*w 105
11M 71
fM f*w = deg(f) -IN w 107
y - p(y) 1- Tp(y)M 71
II = grad (f) / II grad(f) II 108
dist(y,M) = Ily-p(y)11 71
OM = det(lI, .) 108
Symbols 237

dxv : TxM -+ Tx M 108 deg(17(g)) = deg(g) 176


K(x) = det(dxv) 109 f(x) # -g(x) 186
Il,=JMKrlM 109 7r m (M)=Z 193
rlsm 109 7rm(§m) = 7rm(§m) = Z 193
V*rlsm = KrlM 109 deg(f· g) = deg(f) + deg(g) 195
X(M) 109 7rm(x) = Z 198
Il, = ~vol(§m)X(M) 109 §2m-l -+ §m 200
Hl(§l x §l,JR) = JR2 110 H(f 0 g) = H(f) deg(g) 200
deg(f) 111 7r3(§2) =Z 205
f * g: §m+n+l -+ §m+n+l 113 7r7(§4) = Z EB Z4 EB Z3 205
exp: JR -+ §l 113 7r15(§S) = Z EB Zs EB Z3 EB Z5 205
h = - f + (f,g)g : §m -+ §m 113 H(g 0 f) = deg(g)2H(f) 205
deg(f) = deg(1/J 0 f 0 cp-l) 117 ~ -- I: i=l
m ~.~
tax; 206
p, -p,cP,iP 119 e=(6,···'~m) 206
cPa,b 119 dz~ : TzM -+ TzM 207
f-l(a) x f-l(b) -+ §2m-2 119 det(dz~) 207
H(f) = deg(cPa,b) 119 indz(~) = sign det(dz~) 207
£(f-l(a), f-l(b)) 121
deg( nfrr Is) 209
Hm(§2m-l, JR) = 0 123
Ind(~) = I:z indz(~) 209
fl(f) = JS2m-l a /\ da 123
grad (f) 211
deg2(f) 124
Gxcp 212
w2(M,p) 124
Jxe 212
W2(p) 127
Hx(f 0 cp) 212
Ilfll 137
Qz(f) 213
d(f, D, a) 138
Qz(f) = (f 0 'Y)"(t) 213
w(f, a) = dU, D, a) 156
indz(f) 214
f(x) = ±f(-x) 160
Ind(~) = I:z(-l)ind. U ) 214
~#±~
IIf(x)1I IIf(-x)1I 163
Ind(~) = I:z(-l)k ak 214
f(x) _ ~
IIf(x)1I - =i=lIf(-x)1I 166 dist(x, H) = II!II (x, a) 215
limllxll-HOO If(x)1 = +00 169 80(3) 217
s(X), s'(X) 171 Ind(~) = deg(7J) 219
d(gof,D,a) 172 daf = IdE EBda~ 219
17(g) 176 Ind( -~) = (-l)m Ind(~) 220
238 Symbols

Ind(~) = X(M) 220


X=F-E+V 221
X = 2 - 2g 221
X(§2) = 2 221
X(M) = Lk(-l)kak 222
deg(v) = ~X(M) 222
Index
Accessible points, 16 Bump function, 51
Additivity, 37, 39-41, 45, 46, 149
Admissible class of mappings, 41 Canonical orientation, 85
Admissible extension, 44 Cauchy Argument Principle, 5, 6,
Alexandroff compactification, 43 160
Amann-Weiss degree, 42 Cauchy index of a function, 3
Analysis Situs, 13 Cauchy index of a planar curve, 5,
Antipodal extension, 187 11
Antipodal images, 186 Cauchy index with respect to a
Antipodal preimages, 167 planar curve, 15
Approximation and homotopy, 78 Cell, 14
Approximation by smooth Center, 208
mappings, 76 Change of coordinates, 54
Arcs do not disconnect, 15, 180 Change of variables for integrals, 38,
Attractor, 207 107
Axiomatization of Euclidean degree, Characteristic of a regular function
149 system, 8
Circulation, 208
Balls do not disconnect, 180 Classification of compact
Barycentric subdivision, 25 differentiable curves, 63
Bolzano Theorem, 167 Closed differential form, 104
Borsuk Theorem, 195 Cobordism, 42
Borsuk-Ulam Theorem, 163 Cohomotopy group, 75
Boundaries are not retractions, 13, Coincidence index of two mappings,
133 29
Boundary, 14 Colevel of a space with involution,
Boundary of a manifold, 56 171
Boundary of a simplicial complex, Combinatorial topology, 13, 23
23 Completely continuous mapping, 33
Boundary Theorem, 19, 96, 101, Complex multiplication, 100
111, 150, 157 Complex projective space, 61
Brouwer Fixed Point Theorem, 13, Converse Boundary Theorem, 194,
20, 28, 29, 134, 152, 169 199
Brouwer-Kronecker and Euclidean Convolution kernel, 144
degrees, 157, 158 Counterclockwise orientation, 85
Brouwer-Kronecker degree, 31, 99, Critical point, 62
111 Critical value, 35, 62
-239
240 Index

Curve, 54 Euclidean degree is locally constant,


Curve germs, 60, 90 139, 146
Cusp, 60 Euler characteristic, 30, 109, 220
Euler formula, 221
Degree for mappings between spaces Even and odd extension lemmas,
of different dimensions, 43 161
Degree of a complex polynomial, 151 Even and odd mappings, 160
Degree of a complex rational Exact form, 104
function, 190 Excision, 37, 44, 46
Degree of a composite mapping, 27, Exhaustion by compact sets, 54
100, 111 Existence of solutions, 1, 19, 27, 34,
Degree of a differentiable mapping, 37, 40, 44, 46, 149
114,154 Extension by zero, 53
Degree of a symmetry, 100 Exterior differentiation, 104
Degree of the antipodal map, 99 Exterior of a compact hypersurface,
Degree on manifolds with boundary, 129
193 Exterior of a planar curve, 15
Deimling Axiomatization, 149
Deimling axiomatization, 40 Face, 23
Derivative of a differentiable Fibrations of spheres, 202
mapping, 59 Finite representation, 46
Diffeomorphism, 54 First homotopy group, 30
Diffeotopies of Euclidean spaces, 81 Fixed Point Problem, 1
Diffeotopies of manifolds, 82 Fixed Point Theorem for odd
Diffeotopies preserve orientation, 86 mappings, 167
Diffeotopy, 80 Fixed points of mappings of spheres,
Differentiable function, 49 21, 27
Differentiable manifold, 54 Flow of a tangent vector field, 207
Differentiable mappings, 49 Flow through a surface, 11
Differentiable mappings are Framed cobordism, 43
homotopic to their Fredholm operator, 34
derivatives, 153 Fubini Theorem, 65, 216
Differentiable retraction, 67 Fiihrer Axiomatization, 149
Differential form, 104 Fuller index, 45
Dimension of a manifold, 54 Fundamental group, 30
Dipole, 210 Fundamental Theorem of Algebra,
Directional derivation, 58 2, 102, 150
Divergence of a vector field, 105 Fiihrer axiomatization, 39

Easy Sard Theorem, 63, 187 Gauss curvature, 109


Electric field flow, 11, 160 Gauss mapping, 30, 108
Equivariant degree, 46 Gauss Theorem on electric flows, 11,
Equivariant mapping, 46 160
Equivariant mapping of spaces with Gauss-Bonnet Theorem, 31, 109,
involution, 171 222
Equivariant mapping of spheres, 171 Geba-MassabO-Vignoli degree, 43
Equivariant topological degree, 45 Genus of a compact surface, 30, 221
Euclidean degree, 138, 143, 145 Global equations, 55, 129
Index 241

Global normal field, 215 Identity off a set, 80


Good simplicial approximation, 25 Impulsor, 207
Gradient of a smooth function, 58 Incoming (eingang) point, 8
Gradients on arbitrary manifolds, Index of a center, 208
211 Index of a dipole, 210
Gramm matrix, 212 Index of a function at a critical
Green's Theorem, 16 point, 214
Index of a function system on a
Hadamard index, 36 hypersurface, 19
Hadamard Integral Theorem, 17, Index of a non-degenerate zero, 207
159 Index of a quadratic form, 214
Hedge-hog Theorem for manifolds, Index of a saddle, 207
221 Index of a sink, 207
Hedge-hog Theorem for spheres, 135 Index of a source, 207
Height function, 215 Index of an isolated zero, 209
Heinz Integral Formula, 38, 144 Integral curvature, 30, 109
Hessian of a function at a critical Integral index, 3
point, 213 Integral lines, 207
Hirsch Theorem, 166 Interior of a compact hypersurface,
Homology groups, 222 129
Homomorphisms on cohomology, Interior of a planar curve, 15
105 Intermediate Value Theorem in
Homotopic mappings, 75 arbitrary dimension, 12
Homotopy, 75 Invariance of dimension, 28, 169
Homotopy and smooth homotopy, Invariance of Domain Theorem, 28,
79 56, 167, 169, 181
Homotopy for manifolds with Invariance of interiors, 20, 169
boundary, 193 Invariance of the characteristic
Homotopy for non-compact under continuous
manifolds, 193 deformations, 12
Homotopy for non-orientable Invariant set, 46
manifolds, 194 Inverse image of a regular value, 24,
Homotopy group, 30, 75 35, 62
Homotopy invariance, 25, 27, 37, Inverse Mapping Theorem for
39-41, 44, 46, 98, 101, 107, manifolds, 59
112, 122, 140, 142, 144, Inward tangent vectors at a
147, 156 boundary point, 90
Homotopy of mappings into spheres, Isolated coincidence point of two
186 mappings, 29
Hopf fibration, 202, 205
Hopf invariant, 30, 119 Jacobian, 4, 9, 58
Hopf invariant in odd dimension, Join of two mappings, 113
122 Jordan hypersurface, 31
Hopf invariant of a composite Jordan Separation Theorem, 6, 14,
mapping, 200 28, 55, 88, 108, 124, 126,
Hopf Theorem, 29, 191, 196 158, 176, 181
Hypersurface, 54 Jordan sets separate the space, 180
242 Index

Kneser-Glaeser Theorem, 63 Non-degenerate critical point of a


Kronecker Existence Theorem, 9 function, 212
Kronecker index, 14 Non-degenerate zero of a tangent
Kronecker integral, 11-13 vector field, 207
Kronecker Integral Theorem, 10, 159 Non-embedded manifolds, 31
Non-homotopic mappings with the
Lemniscate, 171 same degree, 112
Leray-Schauder degree, 34 Norm of a function, 137
Letter from Brouwer to Hadamard, Normal vector field, 87, 88, 108, 218
21 Normal vector field compatible with
Level of a space with involution, 171 an orientation, 87
Lifting, 113 Normal vector field on a tube, 219
Linearization of antipodal Normality, 27, 37, 39-41, 139, 149
extensions, 188 Normalization of antipodal
Link coefficient, 28, 121 extensions, 188
Liouville-Sturm Theorem, 5
Local coordinate system, 54 Opposite orientations, 85
Local coordinate system compatible Order of the origin with respect to a
with an orientation, 85 surface, 17
Local diffeomorphism, 54 Orientation at two antipodal points
Local differentiable extension, 49 of a sphere, 88
Local equations, 55 Orientation in a linear space, 85
Local extrema of a differentiable Orientation of a cylinder, 91
function, 59 Orientation of a manifold, 85
Local finiteness, 50 Orientation of hypersurfaces, 87
Localization, 24, 59 Orientation of inverse images, 89
Locally closed set, 55 Orientation of spheres, 88
Orientation of the boundary, 91
Manifold, 23 Oriented distance, 215
Manifold with boundary, 56 Orthogonal group, 66
Manifolds are homogeneous, 83 Outgoing (ausgang) point, 8
Manifolds as retractions of open Outward normal vector field, 88
Euclidean sets, 70 Outward tangent vectors at a
Mappings into spheres, 99, 183, 185 boundary point, 90
Mappings into toruses, 186
Measure zero set, 64 Parallelizable manifold, 210
Mod 2 degree theory, 34, 124 Parametrization, 54
Mod 2 winding number, 124 Partial derivative, 58
Models of a manifold, 31 Perron-Fr6benius Theorem, 134
Morse function, 214, 222 Poincare-Bohl Theorem, 19, 159
Morse inequalities, 221 Poincare-Hopf Index Theorem, 21,
Multiplication Formula, 172 30, 109,219
Multiplicities of roots, 151 Poincare-Miranda Theorem, 28
Multiplicities of zeros, 6 Polyhedron, 13
Positive atlas, 85
Nagumo Axiomatization, 149 Positive basis, 85
Nagumo axiomatization, 37 Positive change of coordinates, 85
Noether operator, 34
Index 243

Preserving and reversing orientation Solid torus, 60


mappings, 86 Source, 207
Principal curvatures, 109 Spheres are not homeomorphic to
Product of manifolds, 57 proper subsets, 180
Proper homotopy, 76 Splitting of an isolated zero, 208
Proper mapping, 75 Stereographic projection and
Proper subset of spheres do not orientation, 88
disconnect Euclidean Stereographic projections, 55
spaces, 180 Stiefel bundle, 61, 66
Properly homotopic mappings, 76 Stokes' Theorem, 13, 38, 104
Pseudomanifold, 23 Surface, 54
Pull-back of a differential form, 105 Surface in the Euclidean space, 16
Suspension, 44
Quadratic transform, 144 Suspension of a mapping, 176, 200
Symmetric set, 161
Radial retraction, 61
Real projective hyperplane, 67 Tangent bundle, 61
Real projective hypersurface, 67 Tangent space, 58
Real projective space, 61 Tangent vector, 58
Real roots of a polynomial, 2, 7 Tangent vector field, 135, 206
Regular function system, 7 Tangent vector fields via flows, 221
Regular point, 62 Tangent vector fields via
Regular value, 62 triangulations, 220
Residual set, 63 Tangent vector fields without zeros,
Retraction, 67 21, 27
de Rham cohomology, 37, 104 Tietze Extension Theorem
Riesz Representation Theorem, 211 (differentiable), 52
Rotation group, 217 Topological beast, 16
Topological circle, 183
Saddle, 207 Topological degree, 24
Sard Theorem, 35 Topological degree at the origin, 5
Sard-Brown Theorem, 35, 63 Topological sinus, 113, 145
Sch6nflies Theorem, 16, 20 Torus of dimension m, 102
Screwdriver orientation, 85 Torus with 9 holes, 221
Segre embedding, 189 Total index of a tangent vector field,
Semi-cone, 60 209
Set topology, 15 Total index of the solutions of an
Sign of a mapping at a point, 86 equation in a Banach
Simplex, 23 space, 34
Simplicial approximation, 25 Total number of zeros, 6
Simplicial complex, 23 Transversality, 62
Simplicial homology, 13 Triangulation, 13, 220
Singular cohomology, 104 Thbular retraction, 71
Singularity of a tangent vector field,
206, 212 Unitary normal vector field, 87, 88,
Sink, 207 108, 218
Smooth function, 49 Uryshon separating function, 51
Smoothness, 57, 117
244 Index

Variation of the argument, 14


Vector product, 87
Volume element, 108

Weingarten endomorphism, 108


Whitney Embedding Theorems, 56
Winding number, 5, 156, 163, 166
Winding number is locally constant,
156

Zero of a tangent vector field, 206


Titles in This Series
108 Enrique Outerelo and Jesus M. Ruiz, Mapping degree theory, 2009
107 Jeffrey M. Lee, Manifolds and differential geometry, 2009
106 Robert J. Daverman and Gerard A. Venema, Embeddings in manifolds, 2009
105 Giovanni Leoni, A first course in Sobolev spaces, 2009
104 Paolo Alufli, Algebra: Chapter 0, 2009
103 Branko Griinbaum, Configurations of points and lines, 2009
102 Mark A. Pinsky, Introduction to Fourier analysis and wavelets, 2009
101 Ward Cheney and Will Light, A course in approximation theory, 2009
100 I. Martin Isaacs, Algebra: A graduate course, 2009
99 Gerald Teschl, Mathematical methods in quantum mechanics: With applications to
Schrodinger operators, 2009
98 Alexander I. Bobenko and Yuri B. Suris, Discrete differential geometry: Integrable
structure, 2008
97 David C. Ullrich, Complex made simple, 2008
96 N. V. Krylov, Lectures on elliptic and parabolic equations in Sobolev spaces, 2008
95 Leon A. Takhtajan, Quantum mechanics for mathematicians, 2008
94 James E. Humphreys, Representations of semisimple Lie algebras in the BGG category
0,2008
93 Peter W. Michor, Topics in differential geometry, 2008
92 I. Martin Isaacs, Finite group theory, 2008
91 Louis Halle Rowen, Graduate algebra: Noncommutative view, 2008
90 Larry J. Gerstein, Basic quadratic forms, 2008
89 Anthony Bonato, A course on the web graph, 2008
88 Nathanial P. Brown and Narutaka Ozawa, CO-algebras and finite-dimensional
approximations, 2008
87 Srikanth B. Iyengar, Graham J. Leuschke, Anton Leykin, Claudia Miller, Ezra
Miller, Anurag K. Singh, and Uli Walther, Twenty-four hours of local cohomology,
2007
86 Yulij Ilyashenko and Sergei Yakovenko, Lectures on analytic differential equations,
2007
85 John M. Alongi and Gail S. Nelson, Recurrence and topology, 2007
84 Charalambos D. Aliprantis and Rabee Tourky, Cones and duality, 2007
83 Wolfgang Ebeling, Functions of several complex variables and their singularities
(translated by Philip G. Spain), 2007
82 Serge Alinhac and Patrick Gerard, Pseudo-differential operators and the Nash-Moser
theorem (translated by Stephen S. Wilson), 2007
81 V. V. Prasolov, Elements of homology theory, 2007
80 Davar Khoshnevisan, Probability, 2007
79 William Stein, Modular forms, a computational approach (with an appendix by Paul E.
Gunnells), 2007
78 Harry Dym, Linear algebra in action, 2007
77 Bennett Chow, Peng Lu, and Lei Ni, Hamilton's Ricci flow, 2006
76 Michael E. Taylor, Measure theory and integration, 2006
75 Peter D. Miller, Applied asymptotic analysis, 2006
74 V. V. Prasolov, Elements of combinatorial and differential topology, 2006
73 Louis Halle Rowen, Graduate algebra: Commutative view, 2006
72 R. J. Williams, Introduction the the mathematics of finance, 2006
71 S. P. Novikov and I. A. Taimanov, Modern geometric structures and fields, 2006
TITLES IN THIS SERIES

70 Sean Dineen, Probability theory in finance, 2005


69 Sebastian Montiel and Antonio Ros, Curves and surfaces, 2005
68 Luis Caffarelli and Sandro Salsa, A geometric approach to free boundary problems,
2005
67 T.Y. Lam, Introduction to quadratic forms over fields, 2004
66 Yuli Eidelman, Vitali Milman, and Antonis Tsolomitis, Functional analysis, An
introduction, 2004
65 S. Ramanan, Global calculus, 2004
64 A. A. Kirillov, Lectures on the orbit method, 2004
63 Steven Dale Cutkosky, Resolution of singularities, 2004
62 T. W. Korner, A companion to analysis: A second first and first second course in
analysis, 2004
61 Thomas A. Ivey and J. M. Landsberg, Cartan for beginners: Differential geometry via
moving frames and exterior differential systems, 2003
60 Alberto Candel and Lawrence Conlon, Foliations II, 2003
59 Steven H. Weintraub, Representation theory of finite groups: algebra and arithmetic,
2003
58 Cedric Villani, Topics in optimal transportation, 2003
57 Robert Plato, Concise numerical mathematics, 2003
56 E. B. Vinberg, A course in algebra, 2003
55 C. Herbert Clemens, A scrapbook of complex curve theory, second edition, 2003
54 Alexander Barvinok, A course in convexity, 2002
53 Henryk Iwaniec, Spectral methods of automorphic forms, 2002
52 llka Agricola and Thomas Friedrich, Global analysis: Differential forms in analysis,
geometry and physics, 2002
51 Y. A. Abramovich and C. D. Aliprantis, Problems in operator theory, 2002
50 Y. A. Abramovich and C. D. Aliprantis, An invitation to operator theory, 2002
49 John R. Harper, Secondary cohomology operations, 2002
48 Y. Eliashberg and N. Mishachev, Introduction to the h-principle, 2002
47 A. Yu. Kitaev, A. H. Shen, and M. N. Vyalyi, Classical and quantum computation,
2002
46 Joseph L. Taylor, Several complex variables with connections to algebraic geometry and
Lie groups, 2002
45 Inder K. Rana, An introduction to measure and integration, second edition, 2002
44 Jim Agler and John E. MCCarthy, Pick interpolation and Hilbert function spaces, 2002
43 N. V. Krylov, Introduction to the theory of random processes, 2002
42 Jin Hong and Seok-Jin Kang, Introduction to quantum groups and crystal bases, 2002
41 Georgi V. Smirnov, Introduction to the theory of differential inclusions, 2002
40 Robert E. Greene and Steven G. Krantz, Function theory of one complex variable,
third edition, 2006
39 Larry C. Grove, Classical groups and geometric algebra, 2002
38 Elton P. Hsu, Stochastic analysis on manifolds, 2002
37 Hershel M. Farkas and Irwin Kra, Theta constants, Riemann surfaces and the modular
group, 2001

For a complete list of titles in this series, visit the


AMS Bookstore at www.ams.org/bookstore/.
This textbook treats the classical parts of mapping degree theory, with a detailed
account of its history traced back to the first half of the 18th century. After a
historical first chapter, the remaining four chapters develop the mathematics.
An effort is made to use only elementary methods, resulting in a self-contained
presentation. Even so, the book arrives at some truly outstanding theorems:
the classification of homotopy classes for spheres and the Poincare-Hopf Index
Theorem, as well as the proofs of the original formulations by Cauchy, Poincare,
and others.
Although the mapping degree theory you will discover in this book is a classical
subject, the treatment is refreshing for its simple and direct style. The straight-
forward exposition is accented by the appearance of several uncommon topics:
tubular neighborhoods without metrics, differences between class I and class 2
mappings, Jordan Separation with neither compactness nor cohomology, explicit
constructions of homotopy classes of spheres, and the direct computation of the
Hopf invariant of the first Hopf fibration .
The book is suitable for a one-semester graduate course. There are 180 exer-
cises and problems of different scope and difficulty.

_ For additional information


~ and updates on this book, visit
www.ams.org/bookpages/gsm-I 08

ISBN 978-0-8218-4915-6

9 780821849156
GSM/I08

Potrebbero piacerti anche