Topology, An Invitation
Topology, An Invitation
K. Parthasarathy
Topology
An Invitation
UNITEXT
Volume 134
Editor-in-Chief
Alfio Quarteroni, Politecnico di Milano, Milan, Italy
École Polytechnique Fédérale de Lausanne (EPFL), Lausanne, Switzerland
Series Editors
Luigi Ambrosio, Scuola Normale Superiore, Pisa, Italy
Paolo Biscari, Politecnico di Milano, Milan, Italy
Ciro Ciliberto, Università di Roma “Tor Vergata”, Rome, Italy
Camillo De Lellis, Institute for Advanced Study, Princeton, NJ, USA
Massimiliano Gubinelli, Hausdorff Center for Mathematics, Rheinische
Friedrich-Wilhelms-Universität, Bonn, Germany
Victor Panaretos, Institute of Mathematics, École Polytechnique Fédérale de
Lausanne (EPFL), Lausanne, Switzerland
The UNITEXT - La Matematica per il 3+2 series is designed for undergraduate
and graduate academic courses, and also includes advanced textbooks at a research
level.
Originally released in Italian, the series now publishes textbooks in English
addressed to students in mathematics worldwide.
Some of the most successful books in the series have evolved through several
editions, adapting to the evolution of teaching curricula.
Submissions must include at least 3 sample chapters, a table of contents, and a
preface outlining the aims and scope of the book, how the book fits in with the
current literature, and which courses the book is suitable for.
For any further information, please contact the Editor at Springer: francesca.
[email protected]
THE SERIES IS INDEXED IN SCOPUS
Topology
An Invitation
K. Parthasarathy (Emeritus)
Ramanujan Institute
University of Madras
Chennai, Tamil Nadu, India
Mathematics Subject Classification: 54Axx, 54Bxx, 54Cxx, 54Dxx, 54Exx, 54Fxx, 54Hxx, 55Pxx, 57Nxx
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Singapore Pte Ltd. 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
(Ancient Sanskrit ‘Sloka:’)
One fourth from the teacher, one fourth from
student’s own intelligence,
One fourth from co-learners, and one fourth
in course of time.
Dedicated to
my parents
Lakshmi and Krishnan
during my father’s centenary year
and to
Kannamma
Preface
Mathematics is the most beautiful and the most powerful creation of the human
spirit, said Stefan Banach. One may add: Topology is one of the most beautiful
and powerful branches of mathematics. The word ‘topology’ means ‘study of posi-
tion/place’ and is derived from two Greek words ‘topos’ and ‘logos’. From a math-
ematical point of view, a succinct description of topology may be given as ‘a study
of continuity’. Topology is indispensable in analysis and geometry. Far from being
an abstract subject, it has been growing into a versatile tool of great value in many
branches of science—physics (cosmology, quantum field theory, string theory and
condensed matter physics), biology (molecular structure, membrane biology, neural
networks, evolutionary biology, enzymes and DNA, protein folding), computer
science (topological data analysis, a ‘hot’ subject now), robotics, etc.
This book is intended to be an introduction to topology with historical perspective.
It discusses all the basics of general topology that a student of mathematics would
need and provides glimpses of the beginnings of algebraic topology and combinato-
rial methods in understanding the topology of the sphere and the Euclidean space.
For students having familiarity with elementary real analysis and some felicity with
the language of set theory and abstract mathematical reasoning, this book would
be suitable as a textbook both for a first course in topology and for self-study. The
language of sets is indispensable in modern mathematics, but we do not present or
use any axiomatic set theory, per se. The only exception to this is the use of Axiom of
Choice in the form of Zorn’s lemma in the proof of Tychonoff’s theorem on products
of compact spaces. We give a brief discussion on the Axiom of Choice and some of
its equivalents in that chapter.
This book starts with a discussion of the classical intermediate value theorem
and some of its uncommon consequences as an appetiser to kindle the interest of
the reader before the main course is served. The introduction of general topological
spaces and continuous maps on them in Chaps. 3 and 4 is preceded by a moti-
vating discussion of metric spaces and continuous maps on these (Chap. 2). Several
important special properties of spaces—compactness, connectedness, separation and
countability axioms—are presented in Chaps. 5, 8–10. In between, topologies defined
by maps—initial and final topologies—are presented in Chap. 6. The product and
ix
x Preface
quotient topologies are seen as important special cases of these. A separate chapter
(Chap. 7) is devoted to products of compact spaces. Locally compact spaces are
treated in Chap. 11 and complete metric spaces in Chap. 12. Elementary, combina-
torial treatments of some of the most important results on Euclidean spaces—the
higher-dimensional intermediate value theorem of Poincaré–Miranda, Brouwer’s
fixed point theorem, the no retract theorem, theorems on invariance of domain
and dimension, Borsuk’s antipodal theorem, the Borsuk–Ulam theorem, Lusternik–
Schnirelmann–Borusk theorem—are given in Chap. 13. The last two chapters provide
an introduction to homotopy, fundamental groups and covering spaces.
Although most of the topics discussed in general topology are standard, there
are a few that are not common in topology texts. Instances are: the intermediate
value theorem and its consequences, Bourbaki’s Mittag–Leffler theorem (from which
Baire’s theorem is deduced) and (a simple case of) continuous partitions of unity.
Nets are defined early, after limit points are introduced and are used again later as
needed—for example, to characterise continuity and compactness.
This book has several salient features. An unusual one among them is its historical
perspective. The author’s perception is that learning mathematics at any level should
be spiced up with a dash of historical development. With this in mind, original
formulations for all the basic concepts introduced and original statements of major
results are provided (all with English versions) before modern versions are given.
(The relevant original references, starting from early nineteenth century, are listed in
the bibliography.) This should give some idea of how, and through whom, concepts
developed before they took the current shape. However, this is not a book on the
history of topology—far from it. This is an introductory book on topology that also
wishes to provide a bit of historical flavour of the topics discussed. In addition to
the historical remarks interspersed throughout the text, thumbnail sketches of the
dramatis personae involved in the development of topics discussed are provided at
the end of each chapter. This is intended to enhance the reader’s understanding of
the historical development of the subject.
Poincaré’s higher-dimensional analogue of the intermediate value theorem does
not appear to be well-known even among professional mathematicians (vis-à-vis,
say, the equivalent Brouwer fixed point theorem.) The presentation of elementary
approaches to these and several stellar results like those on domain and dimension
invariance and Borsuk’s results on maps on spheres is another notable feature of the
book.
Recent (one as recent as 2020) and simple proofs of many ‘big’ theorems are
presented. Some instances are: Étienne Matheron’s proof (2020) and a 2003 version
of Paul Chernoff’s proof of Tychonoff’s theorem; Kulpa’s proof of the Poincaré
theorem and Brouwer fixed point theorem following Müger (2015); a 2005 proof
of Ky Fan’s lemma (used in the proof of Borsuk’s theorems); and Kulpa’s proof of
domain invariance theorem (also presented by Terence Tao (2014)).
Numerous exercises, of various levels of difficulty, are given at the end of each
chapter, listed topic-wise for the reader’s convenience. (If the reader is expected to
prove a statement, only the statement is given without the words ‘prove’ or ‘show’.)
Preface xi
References
1. Bourbaki, N.: General Topology, Chapters 1–4. Springer (1995)
2. Bourbaki, N.: General Topology, Chapters 5–8. Springer (1998)
3. Engelking, R.: General Topology Revised Edition. Heldermann Verlag (1989)
4. Kelley, J.L.: General Topology. Affiliated East-West Press (1969)
5. Willard, S.: General Topology. Addison-Wesely (1970)
6. Andre, N., Engdahl, S., Parker, A.: An analysis of the first proofs of the Heine-Borel Theorem-
Borel’s Proof”. Loci (2003). https://doi.org/10.4169/loci003890
1[16] © Mathematical Association of America, 1975. All rights reserved. Quoted here with
permission.
xii Preface
7. Borel, A.: Twenty-Five Years with Nicolas Bourbaki (1949–1973). Not. Am. Math. Soc. 45,
373–380 (1998)
8. James, I.M.: From combinatorial topology to algebraic topology. In: James, I.M. (ed.) History
of Topology, pp. 561–574. Elsevier (1999)
9. Raman-Sundström, M.: A pedagogical history of compactness. Am. Math. Monthly 122,
619–635 (2015). https://doi.org/10.4169/amer.math.monthly.122.7.619
10. Reitberger, H.: The contributions of L. Vietoris and H. Tietze to the foundations of general
topology. In: Aull, C.E., Lowen, R. (eds.) Handbook of the History of General Topology,
vol. 1, pp. 51–72. Kluwer (1997)
11. Reitberger, H.: Leopold Vietoris, 1891–2002. Not. Am. Math. Soc. 49(10), 1232–1236 (2002)
12 Rodriguez, L.: Frigyes Riesz and the emergence of general topology. Arch. History Exact
Sci. 69, 55–102 (2015). https://doi.org/10.1007/s00407-014-0144-6
13. Russ, S.: A translation of Bolzano’s paper on the intermediate value theorem. Historia Math.
7, 156–186 (1980). https://doi.org/10.1016/0315-0860(80)90036-1
14. Sarkaria, K.S.: The topological work of Henri Poincaré. In: James, I.M. (ed.) History of
Topology, pp. 123–167. North-Holland (1999)
15. Thron, W.J. Frederic Riesz’ contributions to the foundations of general topology. In: Aull,
C.E., Lowen, R. (eds.) Handbook of the History of General Topology, vol. 1, pp. 31–40.
Kluwer (1997)
16. Halmos, P.R.: The problem of learning to teach. Am. Math. Monthly 82, 466–476 (1975).
https://doi.org/10.2307/2319737
Contents
xiii
xiv Contents
5 Compact Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.1 Compactness in Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.2 Compactness in Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.3 Compactness in Topological Spaces . . . . . . . . . . . . . . . . . . . . . . . . . 68
5.4 Biographical Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.4.1 E. Borel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.4.2 Heine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.4.3 Lebesgue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.4.4 Cantor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.4.5 Vietoris . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
6 Topologies Defined by Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.1 Initial and Final Topologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
6.2 Product Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
6.3 Quotient Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.4 Biographical Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.4.1 R. L. Moore . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.4.2 Möbius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.4.3 Klein . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7 Products of Compact Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.1 Tychonoff’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.2 Appendix: Axiom of Choice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
7.3 Biographical Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
7.3.1 Bourbaki . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
7.3.2 Čech . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.3.3 Tychonoff . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.3.4 Kelley . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
8 Separation Axioms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
8.1 Hausdorff Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
8.2 Normal Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
8.3 Regular Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
8.4 Completely Regular Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
8.5 Biographical Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.5.1 Tietze . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.5.2 Urysohn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
8.5.3 Carathéodory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
8.5.4 M. H. Stone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
Contents xv
xvii
Chapter 1
Apéritif : The Intermediate Value
Theorem
A function which is continuous either in general, or within given limits, does not
change from one of its values into another value without first having to take all
values lying between them at least once. [1].
That was how Bolzano, a mathematician (and philosopher) of Italian origin based
in Prague, stated his now-famous intermediate value theorem in 1817. The topologi-
cal concept underlying this will be taken up for discussion later. Here, we take a look
at a ‘real analysis’ proof of this theorem and some of its interesting consequences,
including a few that are not often discussed. This is meant as an hors d’oeuvre to
whet the appetite of the reader for the gourmet course to follow.
Bolzano’s result states that a continuous function takes every value between any
two of its values. Using the term ‘interval’ it reads as follows. (Note that a one-point
set is also a (degenerate) interval, so the result holds for constant functions too.)
Theorem 1.2 (Bolzano’s Intermediate Value Theorem) Let I be an interval in R. If
f : I → R is continuous, then the image f (I ) = { f (x) : x ∈ I } is also an interval.
Remark 1.5 Geometrically, this is ‘obvious’: if one end of the graph of f lies below
and the other lies above the x-axis, then it must cut the x-axis at some point in
between. It is remarkable that Bolzano recognised the g.l.b. property of real numbers
decades before Dedekind and Weierstrass, and used it essentially as we have done.
We now discuss some simple but interesting and important consequences. The
higher dimensional analogues will be taken up later.
Corollary 1.7 (Brouwer fixed point theorem, dim. 1) If f : [a, b] → [a, b] is con-
tinuous, then f has a fixed point, i.e. there is a c ∈ [a, b] such that f (c) = c.
We have deduced the fixed point theorem from the intermediate value theorem. Here
is the converse.
Proof Suppose f : [a, b] → R is continuous and f (a), f (b) have opposite signs.
Apply 1.6 to g(x) = x − f (x) if f (a) < 0 and f (b) > 0 and to h(x) = x + f (x)
if f (a) > 0 and f (b) < 0.
(b,b)
y=f(x)
f(c)=c
(a,a)
Proof Assuming that at any instant of time the temperature is a continuous function
of the position, apply Borsuk–Ulam 1.11 for the great circle under consideration in
place of S1 .
Remark 1.13 For this amusing consequence, one may consider, in place of temper-
ature, any other meteorological quantity, e.g. pressure.
Theorem 1.14 (Dosa theorem) Let A1 , A2 be the areas enclosed by two simple
closed curves in the plane. Then there is a straight line which divides both A1 , A2
into two equal parts.
Proof (Following Chinn-Steenrod [2], Shashkin [3]) See Fig. 1.2. We use both the
intermediate value and the Borsuk–Ulam theorems. Enclose both A1 and A2 in a
sufficiently large circle C with centre at the origin. For p on C, let l p be the diametric
line of C through p and L p (x) be the chord perpendicular to l p at a distance x from
p. Then L p (x) divides A1 into two parts, say f 1 (x), g1 (x). If d is the diameter of
C, the intermediate value theorem 1.4 applied to h 1 (x) = f 1 (x) − g1 (x) on [0, d]
yields an x0 in [0, d] such that the chord L p = L p (x0 ) divides A1 into two equal
parts, say f 2 ( p), g2 ( p). Then the function h 2 ( p) = f 2 ( p) − g2 ( p) on C satisfies
h 2 ( p ∗ ) = −h 2 ( p). Apply Borsuk–Ulam 1.11 to h 2 to complete the proof.
Exercise 1
A1 lp
g1 (x)
f1 (x) x0
x A2
p g2 (p)
f2 (p)
Lp (x)
Lp (x0 )
2. Intermediate values. a) If f and g are continuous on [a, b] and if f (a) > g(a)
and f (b) < g(b), then there is a c, a < c < b, such that f (c) = g(c).
b) The function f : [0, 1] → [−1, 1] defined by f (0) = 0 and f (x) = sin(1/x),
x
= 0, is not continuous, yet takes all values in [-1,1].
c) Find a, b such that there is a c, a < c < b, with f (c) = 0, where (i) f (x) =
x 3 − 7x + 23; (ii) f (x) = e−x − x 7 + 17x + sin x − 31.
d) Use the intermediate value theorem to prove that, for any a > 0, there is a
b > 0 such that b2 = a.
e) Apply the intermediate value theorem to conclude that x 4 + x − 3 has at least
two real roots.
f) Use your real analysis to prove Darboux’s theorem: If f is a real, differentiable
function on [0,1], then the derivative f has the intermediate value property.
g) If I is an interval and f : I → R is continuous and injective, then f is mono-
tonic.
3. Fixed points. a) The identity function on [0,1] has every point as a fixed point.
Find continuous functions on [0,1] with (i) exactly 100 fixed points; (ii) countably
infinitely many fixed points; uncountably many fixed points, not the identity function.
b) Find continuous functions X → X without fixed points when X is (i) R; (ii)
[0, ∞); (iii) (0,1]; (iv) the unit circle S1 .
6 1 Apéritif : The Intermediate Value Theorem
4. Dosa theorem. a) If A is a bounded region in the plane, there are two perpen-
dicular lines dividing A into four equal parts.(See [2].)
b) Two dosas, one a perfect circle and the other a perfect square, lie on the table.
Describe a single cut dividing each into two equal parts.
1.2.1 Bolzano
References
1. Bolzano, B.: Rein analytischer Beweis des Lehrsatzes, dass zwischen je zwei Werthen, die
ein entgegengesetztes Resultat gewähren, wenigstens einer reelle Wurzel der Gleichung liege.
Prague, Gottlieb Haase (1817)
2. Chinn, W.G., Steenrod, N.E.: First Concepts of Topology. Mathematical Association of Amer-
ica (1966). https://doi.org/10.5948/UPO9780883859339
3. Shashkin, Y.A.: Fixed Points. American Mathematical Society-Mathematical Association of
America (1991)
Chapter 2
Metric Spaces
The classical ε-δ definition of continuity makes sense once there is a concept of
‘distance’. So, it is natural to begin a study of topology with a look at spaces with a
distance function: metric spaces. The French mathematician Maurice Fréchet defined
a metric in his thesis (1906) [1] thus:
... on peut faire correspondre à tout couple d’éléments A, B un nombre (A, B) ≥ 0
, que nous appellerons l’écart des deux éléments et qui jouit des deux propriété
suivantes: a) L’écart (A, B) n’est nul que si A et B sont identiques, b) Si A, B, C,
sont trois éléments quelconques, on a toujours (A, B) ≤ (A, C) + (C, B). (Fréchet,
1906, [1])
He used the term écart for the ‘distance’ (A, B) ≥ 0 between a pair of elements A
and B (of a set); (a) says (A, B) = 0 if A = B and (b) is the triangle inequality. The
current terminology comes from the German mathematician Hausdorff (see Chap. 3)
who used the term metrische Raumë (‘metric space’). These spaces were vigorously
studied by Hausdorff and the Polish school in the 1920s.
2.1 Metrics
Topology may be called a study of continuity. So let us begin by recalling the concept
of continuity of a function f : R → R; f is continuous at x0 ∈ R if f (x) is arbitrarily
close to f (x0 ) provided x is sufficiently close to x0 . The precise formulation of this
vague, but intuitively appealing, statement results in the famous ε-δ definition of
limits and of continuity. This is attributed to Bolzano (1817, [2]) and, independently,
to the great German mathematician Karl Weierstrass (1872). It is one of the most
fundamental concepts in mathematics.
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 7
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_2
8 2 Metric Spaces
Thus ε measures the closeness of f (x) to f (x0 ) that one is seeking and δ quantifies
the sufficient closeness of x to x0 that is required to achieve this. A rephrasing of this
local formulation is: f (x) is at a distance less than ε from f (x0 ) provided x is at a
distance less than δ from x0 . This makes sense not just on R, but wherever there is a
concept of distance. But what is a distance on a set X ? It clearly involves a pair of
points and is ‘naturally’ nonnegative: it is a map d : X × X → R+ = [0, ∞)! What
kind of a map? Intuition suggests that the distance between distinct points ‘must’ be
positive and distance ‘must’ be symmetric: distance from x to y equals that from y
to x. One also ‘knows’ that the sum of the lengths of two sides of a triangle is greater
than the length of the third side (‘triangle inequality’). Thus the following definition
is fair. It is also mathematically sound and adequate. (‘Metric’ is derived from the
Greek word for measurement.)
Definition 2.2 A metric on a (nonempty) set X is a function d : X × X → R having
the following properties:
i) d(x, x) = 0 and d(x, y) > 0, x, y ∈ X, y = x (Positivity);
ii) d(x, y) = d(y, x), x, y ∈ X (Symmetry);
iii) d(x, y) ≤ d(x, z) + d(z, y), x, y, z ∈ X (Triangle inequality.)
A set X , together with a metric d on it, is called a metric space.
x d(P,y)
d(x,P)
P
which holds for all real numbers λ. For λ = (xi yi )/(yi2 ), this gives
10 2 Metric Spaces
yielding the desired inequality on taking the last term to the left side and then taking
square roots.
It will be convenient to use the Euclidean norm x 2 = d2 (x, 0) for x ∈ Rn . Note that
d2 (x, y) = x − y 2 . Similarly, the norms x 1 := |x j | and x ∞ := max |x j |
give rise to the metrics d1 , d∞ .
Corollary 2.5 (Triangle inequality)
x − y 2 ≤ x 2 + y 2 ,
d2 (x, y) ≤ d2 (x, z) + d2 (z, y) for x, y, z ∈ Rn .
It is time for us to get back to continuity. The ε-δ definition of continuity can now
be carried over to maps between metric spaces.
Proposition 2.7 Let (X, d) be a metric space. For any fixed p ∈ X , the function
x → d(x, p), X → R, is uniformly continuous and d : X × X → R is continuous.
ε
δ f (x0 )
f
x0
Continuity can be rephrased in terms of balls; in metric spaces, balls play a role
similar to that played by intervals in R.
Definition 2.8 Let (X, d) be a metric space. For x ∈ X and r > 0, write B(x; r ) =
{y ∈ X : d(y, x) < r }. It is called the open ball of radius r centred at x. When
necessary, we write B X (x; r ) or Bd (x, r ) for this ball. B(x; ε) is also referred to
as the ε-neighbourhood of x. The closed ball is defined by B̄(x; r ) = {y ∈ X :
d(y, x) ≤ r }. E ⊂ X is called a neighbourhood of x ∈ X if B(x, ε) ⊂ E for some
ε > 0 (Figs. 2.3, 2.4 and 2.5).
Thus, in terms of balls, the defining condition for continuity of a map f : X → X
d∞
d2
d1
Z3
Z2 Q=(1,0)
P=(0,0)
Z1
d1
d2
d∞
i) x ∈ B X (x0 ; δ) ⇒ f (x) ∈ B X
( f (x0 ); ε);
ii) B X (x0 ; δ) ⊂ f −1 (B X
( f (x0 ); ε));
iii) f (B X (x0 ; δ)) ⊂ B X
( f (x0 ); ε).
Example 2.9 1. In R, a ball is an interval: B(x; ε) = (x − ε, x + ε) and B̄(x; ε) =
[x − ε, x + ε].
2. In (R2 , d2 ), B(x; ε) is the open disc with centre x and radius ε.
3. In (R2 , d1 ), B(x; ε) is the open square centred at x with diagonals of lengths
2ε parallel to the axes.
4. In (R2 , d∞ ), B(x; ε) = (x − ε, x + ε) × (x − ε, x + ε) is the open square with
centre x and of sides of length 2ε parallel to the axes.
5. If X has the discrete metric, then B(x, r ) = {x} if r < 1 and B(x, r ) = X if
r ≥ 1.
6. If E ⊂ X , then B E (x; ε) = B X (x; ε) ∩ E.
Next we turn to formulation of continuity in terms of convergence. Sequential
convergence in metric spaces is defined exactly as in R.
Definition 2.10 A sequence {xn } in a metric space (X, d) is said to converge to
x ∈ X , written xn → x, if the sequence {d(xn , x)} of real numbers converges to
2.2 Continuity and Open Sets 13
zero; i.e. if for every ε > 0 there is a positive integer n ε such that d(xn , x) < ε for all
n ≥ n ε . This is expressed by saying that the xn eventually lie in any neighbourhood
of x.
Again, as in the case of the real line, if a sequence in a metric space has a limit,
it is unique.
Proposition 2.11 Let X be a metric space and let {xn } be a sequence in X . If xn → x
and xn → y, then x = y.
Proof If x = y, then ε = d(x, y) > 0 and B(x; ε/2) and B(y; ε/2) are disjoint, so
the xn cannot lie eventually in both.
Example 2.12 1. Convergence in Rn with respect to any of the metrics d1 , d2 , d∞
is equivalent to coordinatewise convergence: x (k) = (x1(k) , ..., xn(k) ) converges to x =
(x1 , ..., xn ) if and only if x (k)
j → x j for all j.
2. In the discrete metric d on a set X , the only convergent sequences are the
eventually constant ones: if xn → x then xn = x for all n ≥ n 0 , for some positive
integer n 0 and conversely.
Proposition 2.13 Let X, X
be metric spaces and let x0 ∈ X . For a map f : X →
X
, the following conditions are equivalent:
1) f is continuous at x0 ;
2) the sequence { f (xn )} converges to f (x0 ) in X
for any sequence {xn } in X
converging to x0 , i.e. d(xn , x0 ) → 0 ⇒ d
( f (xn ), f (x0 )) → 0.
Proof Assume f is continuous at x0 . Let xn → x0 and ε > 0. By continuity, choose
δ > 0 such that d
( f (xn ), f (x0 )) < ε when d(xn , x0 ) < δ and then by convergence
choose n δ such that d(xn , x0 ) < δ for all n ≥ n δ . Then d
( f (xn ), f (x0 )) < ε for
n ≥ n δ , so f (xn ) → f (x0 ).
Conversely, suppose f is not continuous at x0 . This means there is an ε > 0
for which no δ ‘works’; that is, for every δ > 0 there is an x with d(x, x0 ) < δ,
but d
( f (x), f (x0 )) ≥ ε. (The reader should understand this clearly and care-
fully. Forming negations of statements is an indispensable skill in mathematics.)
Thus, for each positive integer n, there is an xn ∈ X such that d(xn , x0 ) < n1 but
d
( f (xn ), f (x0 )) ≥ ε. But then xn → x0 in X , whereas the f (xn ) lie away from
f (x0 ) and so cannot converge to f (x0 ) in X
. This completes the proof.
Remark 2.14 The statement ‘if P is true, then Q is true’ is equivalent to the statement
‘if Q is false, then P is false’. The latter statement is called the contrapositive of the
former. Using contrapositive is an important and frequently employed strategy in
mathematical proofs.
Example 2.15 1. Let X be a metric space and let f : X → Rn . Write f (x) =
( f 1 (x), ..., f n (x)), x ∈ X . Then f is continuous at a ∈ X if and only if each coordi-
nate function f j : X → R is continuous at a.
2. For E ⊂ X , f (x) = d(x, E) := inf y∈E d(x, y) is continuous on X .
3. Any map from a discrete metric space X to any Y is continuous.
14 2 Metric Spaces
ε
y
d(x,y) r
x
Proposition 2.17 In a metric space X , the whole set X , the empty set and any ball
B(x; r ) for x ∈ X and r > 0 are open.
Proof The definition applies trivially to X and vacuously to the empty set. To prove
that B(x; r ) is open, let y ∈ B(x; r ). To find an ε > 0 such that B(y; ε) ⊂ B(x; r ).
Take ε ≤ r − d(x, y). If d(z, y) < ε then
Example 2.18 1. Let a, b ∈ R, a < b. The intervals (a, b), (a, ∞), (−∞, a) are all
open sets in R with the usual metric. Note that (a, b) = (c − r, c + r ) = B(c; r ), c =
a+b
2
and r = b−a 2
. The reader should check the assertion for the other two intervals.
Other than R itself, these are the only intervals, bounded or unbounded, that are open
sets in R. Thus, the only intervals that are open sets are the ‘open’ intervals, bounded
or unbounded.
2. The half planes {(x, y) ∈ R2 : x < a}, {(x, y) ∈ R2 : y < a} and similarly
{(x, y) ∈ R2 : x > a}, {(x, y) ∈ R2 : y > a} are all open in R2 with any of the met-
rics d1 , d2 or d∞ . So are the vertical and horizontal strips {(x, y) ∈ R2 : a < x < b}
and {(x, y) ∈ R2 : c < y < d}.
The class of open sets in a metric space has the following very basic properties
with respect to unions and intersections.
A consequence is that all open sets can be built up from open balls.
Corollary 2.22 A subset V of a metric space X is open if and only if it is a union of
open balls.
Example 2.23 Infinite intersections of open sets are not generally open. For instance,
look at the sequence of open intervals Un = {(− n1 , n1 )}∞
1 in R. Then ∩n Un = {0}
which certainly is not open.
Open sets in R are unions of open intervals, by Corollary 2.21. But an impressively
stronger result holds.
Theorem 2.24 Every nonempty open set in R is a countable, disjoint union of open
intervals.
16 2 Metric Spaces
Proof Let V ⊂ R be open. For each x ∈ V , let Ix be the union of all open intervals
containing x and contained in V :
Ix = {I : I is an open interval and x ∈ I ⊂ V }.
Note that the collection on the right side is nonempty since V is open. Moreover,
Ix is an interval (being a union of intervals with a common point), is open and is
a subset of V . This ensures that if z ∈ Ix , then Ix ⊂ Iz ⊂ Ix so that Iz = Ix . Thus,
if x, y ∈ V and if z ∈ Ix ∩ I y , then Ix = Iz = I y . In other words, if x, y ∈ V , then
Ix , I y are either disjoint or equal. Clearly, the union of the Ix , x ∈ V, is V itself. It
remains to prove countability. For this we use rationals (what else!). Since every open
interval contains rationals, there is a rational number qx ∈ Ix for x ∈ V . If Ix = I y ,
then Ix and I y are disjoint and so qx = q y . Thus Ix → qx is one-one map of the set
of distinct Ix into Q. Hence this collection is countable and the proof is complete.
The property of Q that we have used (‘every open interval contains a q ∈ Q’)
is shared by other countable
√ sets too, e.g. the set of dyadic rationals (those of the
form a/2k ) and {m + n 2 : m, n ∈ Z} (Proof?). Such sets, ‘dense sets’, would be
discussed later in a general setting.
As we observed earlier, convergence in all the metrics d1 , d2 , d∞ on Rn are the
same (= coordinate-wise convergence). This reflects the interesting fact that all of
them yield the same class of open sets.
Proof The equivalences are consequences of (1) and (2) of the previous lemma,
respectively. The crux of the matter is that each di -ball centred at any point x contains
√ If V is d2 -open and x ∈ V ,
a d j -ball centred at x for all i, j. Here is a sample argument.
then Bd2 (x; ε) ⊂ V for some ε > 0. But d2 (x, y) ≤ n d∞ (x, y), so Bd∞ (x; √εn ) ⊂
Bd2 (x; ε) ⊂ V and V is d∞ -open. The proofs for the remaining parts are on similar
lines, and the reader should have no difficulty with them.
Exercise 2
1. The space 2 . Let 2 = {x = (xn ) : x||2 = |xn |2 < ∞}. d2 , defined as d(x, y)
= x − y , just as in Rn .
2. Metrics. a) On (0,1), d(x, y) =| x1 − 1y | defines a metric.
b) If d is a metric on X , then the associated bounded metric d ∗ and d (x, y) =
d(x,y)
1+d(x,y)
are both metrics and are equivalent to d. More generally, d f (x, y) =
f (d(x, y)) gives a metric for any nondecreasing f : [0, ∞) → [0, ∞) satisfying
f (x + y) ≤ f (x) + f (y) and f (x) = 0 if and only if x = 0.
c) A finite sum ak dk is a metric if each dk is and ak > 0.
d) If f is a continuous real function a metric space (X, d), then d f (x, y) =
d(x, y) + | f (x) − f (y)| is a compatible metric.
1
e) d1 ( f, g) = 0 | f (t) − g(t) | dt, dsup ( f, g) = sup0≤x≤1 | f (x) − g(x)| both
give metrics on C[0, 1]; they are not equivalent.
f) Take the closed unit disc with the usual metric. Write down explicitly the French
railway metric (2.3) taking P as the origin.
g) For 0 < p < 1, let p be the space of all real sequences (xn ) with |xn | p < ∞.
d((xn ), (yn )) = |xn − yn | p gives a metric on this space.
3. Unions of intervals. a) The union of a collection of intervals in R having a
common point is an interval. (An alert reader might have noticed that this fact was
used in the proof of Theorem 2.24.)
b) The decomposition in 2.24 is unique.
4. Convergence. Discuss convergence of sequences in the French railway metric
and in a product of metric spaces.
5. Isometries. A map f of a metric space X to itself is called an isometry if
d( f (x), f (y)) = d(x, y) for all x, y ∈ X .
a) Translations Ta (x) = x + a, a ∈ Rn , and orthogonal transformations, ρ A (x) =
Ax, A an orthogonal matrix, are isometries on Rn .
18 2 Metric Spaces
2.3.1 Fréchet
References
1. Fréchet, M.: Sur quelques point du calcul fonctionnel. Rendiconti di Palermo 22, 1–74 (1906).
https://doi.org/10.1007/BF03018603
2. Bolzano, B.: Rein analytischer Beweis des Lehrsatzes, dass zwischen je zwei Werthen, die
ein entgegengesetztes Resultat gewähren, wenigstens einer reelle Wurzel der Gleichung liege.
Prague, Gottlieb Haase (1817)
Chapter 3
Topological Spaces
We have seen that continuity in metric spaces can be expressed wholly in terms of
open sets and that the class of open sets satisfies some simple set theoretic properties.
This led to an abstract study of a set and classes of subsets with these properties so that
continuity can be defined in this setting, giving rise to general topological spaces. It
is generally accepted that Riemann was the originator of topology. He, in his work on
function theory, algebraic functions and in his famous inaugural lecture on geometry
(Über die Hypothesen welche der Geometrie zu Grunde liegen, [1]) had expressed
ideas on formulating the concept of topological spaces, subspaces of such spaces and
a general theory of such spaces. These may be considered as precursors of modern
topology.
The genesis of topology in the seminal work of Riemann was followed by a huge
body of work of several great mathematicians. Cantor’s set theory and theory of
derived sets, Poincaré’s epochal work on Analysis situs, Fréchet’s metric spaces,
F.Riesz’s study of abstract spaces using a version of limit points and Weyl’s study [2]
of abstract spaces in terms of neighbourhood systems finally culminated in the work
of Hausdorff. Hausdorff’s work ushered in what is now called general topology. Here
is the definition of a topological space from Hausdorff [3]:
Unter einem topologischen Raum verstenhenwir eine Menge E, worin den Ele-
menten (Punkten) x gewisse Teimengen Ux zugeordnet sind, die wir Umgebungen
von x nennen, und zwar nach Massgabe der folgenden Umgebungsaxiome ( A topo-
logical space is a set E together with sets Ux assigned to each x ∈ E satisfying the
following neighbourhood axioms):
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 21
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_3
22 3 Topological Spaces
Example 3.2 1. The class of open sets in a semi-metric or a metric space X form a
topology on X , the topology induced by the semi-metric or metric. This has been the
motivation for our definition of a topology. A topology that is induced by a metric is
called a metric topology. Such a topological space is said to be metrisable.
2. A trivial and hardly useful example is the indiscrete topology on a nonempty
set X consisting of just the whole set and the empty set.
3. A trivial looking, but pretty useful, topology is the discrete topology on X
consisting of all subsets of X .
4. The topology on Rn given by the metric d2 (or, equivalently, by d1 or d∞ ) is
called the Euclidean or the usual topology.
5. If T is a topology on X , for any E ⊂ X , the collection of sets T E := {E ∩ V :
V ∈ T } is a topology on E known as the relative topology. E with this topology is
referred to as a subspace of X . For example, [0,1], [0,1), (0,1], (a, b), 0 < a < b < 1,
are open in [0,1].
One can list a number of examples whose utility is primarily confined to providing
examples and counter examples in specific situations involving various topological
properties. Here are a few of them.
6. On any X, Tco f consists of the empty set and any set V ⊂ X whose complement
V c is finite. It is a topology known as the cofinite topology. The cocountable topology
Tcoc is defined analogously, replacing ‘finite’ by ‘countable’. The cofinite topology
is nondiscrete when X is infinite and the cocountable topology is discrete unless X
is uncountable. Do you get a topology if ‘countable’ is replaced by ‘infinite’?
7. The set X = {0, 1} with the topology {X, ∅, {1}} is called the Sierpiński space S.
It will be useful for counter examples. {X, ∅, {0}} is ‘essentially the same’ topology.
The discrete and the indiscrete topologies are the only other topologies on X .
8. If E is a nonempty, proper subset of a set X , then {X, ∅, E} is a topology on X
and so is {X, ∅, E, E c }.
9. The collection {V ⊂ X : V = ∅ or x0 ∈ V }, for any fixed point x0 ∈ X , is a
topology on X .
10. Here is an interesting topology on Z. For a, b ∈ Z, b > 0, the set A(a, b) :=
{a + nb : n ∈ Z} is a doubly infinite arithmetic progression around a. Define a subset
V of Z to be open if it is empty or satisfies the following property: for every a ∈ V
there is a positive integer b such that A(a, b) ⊂ V . The reader is urged to carry out the
simple verification that this actually gives a nondiscrete topology Tap , the arithmetic
progression topology, on Z. (For a slightly different connected topology (a term to
be defined in Chap. 9) on Z, see [7].)
Sometimes it is useful to compare different topologies on the same set. The fol-
lowing definition will facilitate handling such situations.
Definition 3.6 If T1 and T2 are both topologies on a set X , we say that T1 is finer
or stronger than T2 if T1 ⊃ T2 ; in this case we also say that T2 is coarser or weaker
than T1 .
Example 3.7 On any set, the discrete topology is the finest and the indiscrete topol-
ogy is the weakest topology. The cofinite topology is weaker than the cocountable
topology. On the real line, the usual topology is stronger than the cofinite topology,
but is not comparable with the cocountable topology (i.e. neither is finer than the
other).
Remark 3.8 Thus the family of all topologies on a nonempty set is a partially ordered
set (with inclusion as the order relation) in which any two members have a least upper
bound (lub or sup) and a greatest lower bound (glb or inf ). Such a partially ordered
set is called a lattice. The lattice of all topologies on a set has a smallest element (the
indiscrete topology) and a largest element (the discrete topology).
To conclude the section, we observe that any subset of a topological space contains
a largest open subset.
The next result summarises the properties of the interior, all of which are almost
obvious from the definition.
3.1 Topologies and Open Sets 25
Proof Being a union of open sets, int E is open and (3) is a consequence of (1). All
the other statements are immediate.
Example 3.11 1. If X has the discrete topology, int E = E for E ⊂ X ; if X has the
indiscrete topology, then int E = ∅ if E = X .
2. In R, int Q = ∅ = int (R \ Q) and int [0,1] = (0,1) = int [0,1]∪Z.
3. In the space [0,1] (that is, [0,1] with the relative topology as a subspace of R),
int [0,1] = [0,1], int [0,1) = [0,1) and int (0,1) = (0,1).
4. {(x, 0) : x ∈ R}, {(x, y) : x ∈ Q, y ∈ R} and {(x, y) : y = x 2 } all have empty
interiors in R2 ; the interior of the set {(x, y) : y ≥ 0} is {(x, y) : y > 0}, {(x, y) :
x 2 + y 2 ≤ 1} has interior {(x, y) : x 2 + y 2 < 1}.
5. Consider X = {1, 2, 3} with topology T = {X, ∅, {1}, {2, 3}}. Then int {1} =
{1} = int {1, 2} = int {1, 3} and int {2} = ∅ = int {3}.
Proof All the properties are almost immediate. Property (1) is obvious. If U, V ∈
Ux , then x ∈ int U ∩ int V = int (U ∩ V ), so U ∩ V is in Ux , proving (2). The fact
that U ⊂ V implies int U ⊂ int V gives (3). For (4), just take V = int U .
26 3 Topological Spaces
Open intervals in the real line are basic open sets in the sense that these are the
building blocks of all open sets: every open set in the real line is a union of open
intervals. In metric spaces open balls play an analogous role. Such possible building
blocks on general topological spaces would obviously be of interest and importance.
Definition 3.13 Let X be a topological space. An open base, or just a base or basis,
for (the topology on) X is a collection B of open sets such that every open set in X
is the union of members of some subcollection of B. Each B ∈ B is called a basic
open set.
A family Bx of neighbourhoods of x is a called a neighbourhood base or a local
base at x if every neighbourhood of x contains some B ∈ Bx . Sets in Bx are referred
to as basic neighbourhoods of x.
Proof Part (1) just rephrases the fact that X = ∪{B : B ∈ B}.
(2) If B1 , B2 ∈ B then B1 ∩ B2 is open and so x ∈ B1 ∩ B2 implies there is a B ∈ B
with x ∈ B ⊂ B1 ∩ B2 (Fig. 3.1).
Now let us see if the process can be reversed. In other words, we ask: If we
start with a collection B of subsets of a set X , is there a topology on X for which
B is a base? Such a B necessarily has to satisfy conditions (1) and (2) above. It is
a pleasant fact that these conditions are also sufficient for B to form a base for a
(unique) topology on X .
B
B1 B2
x
U2 × V2
U1 × V1
(Fig. 3.2), it is easy to see that B satisfies the requisite conditions to be a base
for a topology on the product set X × X , called the product topology on X × X . It
will be taken up for a detailed study later.
6. Arithmetic progressions A(a, b) form a base for the topology for Tap (3.2) on
Z.
7. In a metric space X , open balls form a base and open balls B(x, n1 ), n ∈ N,
form a local base at x ∈ X .
8. The collection of unbounded open intervals, i.e. intervals of the form (a, ∞)
and (−∞, b), a, b ∈ R, is not a base for any topology on R, but finite intersections
of these give all open intervals which form a base for the usual topology. To take
care of situations like these, we give the following useful definition (first introduced
by Bourbaki).
R2 .
4. For a set X , spaces Yi and a family of maps f i : X → Yi , the collection S of
all sets of the form f i−1 (Vi ), Vi ⊂ Yi open, is a subbase for a topology on X , known
as the initial topology defined by the f i . Such topologies will be studied in detail in
Chap. 6.
Closed sets are complements of open sets, but the notions of ‘open sets’ and ‘closed
sets’ are not complementary!
3.3 Closed Sets 29
Proposition 3.21 In any topological space X , X and ∅ are closed and arbitrary
intersections and finite unions of closed sets are closed.
Proof Immediate from the definition. Look at unions and intersections of open sets
and take complements.
Infinite unions of closed sets may not be closed: ∪n∈N [1/n, 1] = (0, 1] is not
closed. A family {E i } of subsets in a space X is locally finite if each x ∈ X has a
neighbourhood Ux which meets only finitely many E i .
Proposition 3.22 The union of a locally finite family of closed sets is closed.
Example 3.23 1. In a discrete space every subset is closed and in an indiscrete space
there are no nontrivial closed sets.
2. [0,1] is closed in R (being the complement of (−∞, 0) ∪ (1, ∞)) and so are
Z, {0}; Q is not. {[n, n + 1]}∞ 0 is a locally finite family of closed sets whose union
[0, ∞) is closed.
3. In R2 , the sets {(x, 0) : x ∈ R} and {(x, y) : x 2 + y 2 = 1} are both closed;
Q × Z is not.
4. In a metric space X , single point sets {x} (hence finite sets) are closed: if y = x,
then B(y; d(x, y)) ⊂ X \ {x}, so the last set is open.
5. Every closed ball B̄(x; r ) in any metric space X is a closed set. Reason: if z ∈
X \ B̄(x; r ), then d(z, x) > r and B(z, ε) ⊂ X \ B̄(x; r ) for any 0 < ε ≤ d(z, x) −
r (check); thus B̄(x; r )c is open.
6. In the Sierpiński space S (3.2), {0} is closed, {1} is not.
7. Consider R (3.17). For real numbers a, b, (−∞, a) = ∪∞ 1 [−n, a) is open and
so is [b, ∞) = ∪∞ 1 [b, n). Thus [a, b), a < b, is closed. (Recall that such intervals
form basic open sets for R . So, in R , the intervals [a, b) are open sets which are
also closed.)
Recall that every set has a largest open subset (namely, the interior of the set).
Complementarily, every set has a smallest closed superset.
Definition
3.24 If A is a subset of a space X , the closure Ā of A is defined by
Ā = {F ⊂ X : F ⊃ A, F is closed}.
Proof (1) is a consequence of the definition (the reader should explicitly spell out
the arguments) and (2) is immediate from (1). To prove (3), observe that if x ∈ / Ā,
then x ∈
/ F for some closed set F ⊃ A and V = F c is a neighbourhood of x disjoint
from A. This proves one part of (3). For the converse, retrace the steps to reverse the
arguments. Finally, (4) follows from (2) and the fact that Ā is closed.
The class of closed sets determine the topology and closed sets are precisely those
sets which are fixed under the closure operation A → Ā, so one can expect that the
closure operation determines the topology. This is, indeed, the case; see the exercises
for a precise formulation.
Example 3.26 1. In R, the closure of (0,1) is [0,1] and the closures of Q and R \ Q
are both R. (Recall the construction of real numbers from rational numbers. What
are the properties of rationals and irrationals that are used to get the last statement?)
2. In R (3.17), the closure of (0,1) is [0,1) and the closures of Q and R \ Q are
both R .
3. Consider R with the cocountable topology. The closure of (0,1) is the whole of
R, the closure of Q is itself and that of R \ Q is R.
4. In Rn with the Euclidean metric, the closure of an open ball B(x; r ) is the
corresponding closed ball B̄(x; r ). (Warning: the assertion is not true in a general
metric space!)
5. In a discrete space Ā = A for every set A and in an indiscrete space Ā = X
unless A is empty.
Proof If x ∈ Ā, then A ∩ B(x; n1 ) is nonempty for each positive integer n, so there is
an xn ∈ A ∩ B(x; n1 ) and xn → x. Conversely, if x ∈ / Ā, then there is a ball B(x; ε)
disjoint from A, that is d(y, x) ≥ ε for all y ∈ A, and no sequence in A can converge
to x.
Limits of sequences from a closed set are always in the set. But a set with this
property may not be closed in a topological space, as the following example shows.
Example 3.31 Consider the space R with the cocountable topology. The only closed
sets different from the whole space are countable. So if I = R \ Q is the set of
irrational numbers, then Ī = R. If {xn } is a sequence in R with xn → x, then V =
{x} ∪ R \ {xn : n = 1, 2, ...} is a neighbourhood of x and so must contain the xn for
all but finitely many n. Clearly, this is possible only when the sequence is eventually
the constant x. In other words, the only convergent sequences are those which are
eventually constant. In particular no rational number is a limit of a sequence of
irrationals. Thus, the previous proposition fails in this case. We conclude that for
arbitrary topological spaces, sequences are inadequate to describe the topology.
Clearly, any closed set is sequentially closed, so open sets are sequentially open.
The previous example shows that the set of irrationals is sequentially closed but
not closed in the cocountable topology of R and so the reverse implication is not
generally true. But in metric spaces, the class of closed sets is the same as the class
of sequentially closed sets, as we have seen earlier (3.29).
Proposition 3.33 Let (X, T ) be a topological space and let Tseq be the collection
of all sequentially open sets in X . Then Tseq is a topology on X that is stronger than
T.
Proof The only property that requires verification is that the intersection of two
sequentially open sets is sequentially open, or, equivalently, that the union of two
sequentially closed sets is sequentially closed. So, let E, F be sequentially closed in
X and let A = E ∪ F. Suppose A is not sequentially closed. Then there is a sequence
32 3 Topological Spaces
Proof Suppose D intersects every nonempty open set. Then for any x ∈ X , any
neighbourhood V of x intersects with V and so x ∈ D̄. Thus D̄ = X . On the other
hand, if there is a nonempty open set V disjoint from D, then x ∈
/ D̄ for x ∈ V and
hence D is not dense.
Example 3.36 1. Q, the set of rationals, and R \ Q, the set of irrationals are both
dense in R.
2. In an indiscrete space, every nonempty set is dense and in a discrete space no
proper subset is dense.
3. In the Sierpiński space S (3.2), {1} is dense, but {0} is not.
The next two examples were mentioned in passing earlier.
4. Here is a ‘small’ set of rationals that is also dense in R. A rational number of the
form m/2n , m, n ∈ Z is called a dyadic rational. It is an often used fact that the set
of dyadic rationals is dense in R. Proof : a < b implies 2n (b − a) > 1 for large n, so
the interval (a.2n , b.2n ) has length greater than one and hence contains an integer m:
a.2n < m < b.2n or a < m/2n < b. In other words, every interval contains a dyadic
rational. Hence the assertion. √
5. Now a ‘small’ dense set of irrationals: {m + n 2 : m, n ∈ Z} is a (countable)
dense set in R. This is an exercise for the reader.
Exercise 3
b) Describe the topology on R2 having as a subbase i) all straight lines; ii) all
straight lines parallel to x-axis.
7. Order topology. Ordered spaces were first defined and studied by Eilenberg,
[8]. Let X be a totally ordered set and define intervals as in the real line. The order
topology has as a subbase all sets of the form L a := {x ∈ X : a < x} and Ub := {x ∈
X : x < b}, a, b ∈ X .
a) On R, the usual order gives the usual topology.
b) Write down a base for the order topology on X . (Warning: X may have a
smallest element or a largest element or both or neither.)
c) On the product X × Y of ordered sets, (x, y) < (x , y ) if x < y or x = y and
x < y gives a total order, called the dictionary order. The order topology on N is
the discrete topology, but the order topology on N × N with the dictionary order is
not the discrete topology.
d) What is the order topology on R × R with the dictionary order?
e) In a space X with order topology, is [a, b] the closure of (a, b)?
f) On [0, 1] × [0, 1] with the dictionary order topology, describe neighbourhoods
of (0,0) and (1,1).
g) Compare the dictionary order topology on [0, 1] × [0, 1] with the subspace
topology of [0, 1] × [0, 1] as a subspace of R × R with the dictionary order topology.
h) The order topology induced by the induced order may not be the same as the
relative order topology. Make this precise.
8. Subspaces. a) A subspace of a subspace is a subspace, that is, if A ⊂ B ⊂ X ,
then the relative topology on A as a subspace of X is the same as its relative topology
as a subspace of the subspace B of X .
b) Let X be a space and A ⊂ E ⊂ X . Discuss the relation between:
i) interior of A in E and interior of A in X ;
ii) closure of A in E and closure of A in X .
c) Let E be a subspace of X . Does a basis (or a subbasis) of X give rise to a basis (or
a subbasis) for E? For example, if B is a basis for X , is E ∩ B := {E ∩ B : B ∈ B}
a basis for E?
9. Topology from neighbourhood bases. a) Obtain a result on comparing two
topologies in terms of local bases, analogous to (3.16).
b) Let X be a set and suppose a collection Ux of subsets of X is assigned
for each x ∈ X satisfying the properties (i)–(iv) listed in Proposition 3.12. Define
T = {U ⊂ X : x ∈ U ⇒ U ∈ Ux }. Then T is a topology on X and Ux is precisely
the neighbourhood system at each x ∈ X in this topology. Obtain an analogous char-
acterisation for a collection Bx to form a neighbourhood base at x ∈ X .
c) Fix a prime p. For n ∈ Z and k ∈ N, let Unk = {n + mp k : m ∈ Z}. Then Un =
{Unk : k ∈ N} is a local base at n ∈ Z for a topology on Z.
10. Closures and interiors. a) Examine the validity of the following:
i) the closed ball B̄(x, r ) (in a metric space) is a closed set;
ii) the interior of B̄(x, r ) is the open ball B(x, r );
iii) the closure B(x.r ) of B(x, r ) is the closed ball B̄(x.r ).
b) A ∪ B = Ā ∪ B̄. What about infinite unions? Intersections?
3.3 Closed Sets 35
c) If E is a subset of the real line which is bounded above, then sup E ∈ Ē. Is it
true that sup E is a limit point of E?
d) Find the closure of the graph of f (x) = sin 1/x, 0 < x ≤ 1.
e) True or false? i) Ē = int E; ii) int E=int Ē;
iii) ( Ē)c = int E c ; iv) (int E)c =E c .
f) If ∪ Āi is closed, then ∪ Āi = ∪Ai .
g) If (X, d) is a metric space and A, B ⊂ X , then diam Ā= diam A and d( Ā, B̄) =
d(A, B).
h) Any infinite closed subset of R is the closure of a countable set. Does the result
hold for any Rn ?
i) What is the closure of {1/n : n ∈ N} in R with the cofinite topology? cocount-
able topology?
j) In R (3.17), x ∈ Ā√ if and√
only if there is√
a sequence
√ xn ≥ x with |xn − x| → 0.
Find the closures of (1, 2), ( 2, 2) and (− 2, 2).
k) Let X be the subspace of R2 consisting of the unit circle together with the
diameter on the x-axis. For p = (0, 0), determine the open ball B X ( p, 1), its closure
and the closed ball B̄ X ( p, 1).
11. Closure of the interior vs interior of the closure. (See Bourbaki, [9].) If
X is a topological space and A ⊂ X , let α(A) =int Ā and β(A) = int A.
a) A ⊂ α(A) if A is open and β(A) ⊂ A if A is closed.
b) α(α(A)) = α(A) and β(β(A)) = β(A) for any A.
c) If U and V are disjoint open sets, then so are α(U ) and α(V ).
d) Find a subset A of the real line such that A, int A, Ā, α(A), β(A), β( Ā) and
α(int A) are all distinct.
12. Dense sets. a) D is dense if and only if D c has empty interior.
b) If A, B are dense subsets, A ∩ B may not be dense. (In fact, they can be
disjoint.) However if A, B are also open, then A ∩ B is dense.
c) If D is dense in X , is D ∩ E dense in E for any E? open E?
d) If a subset D intersects all sets in a subbasis, is D necessarily dense? What
happens if it intersects all sets in a basis?
e) Find an explicit countable dense subset in R \ Q.
f) If D is dense in A then it is dense in Ā.
g) When does a subset meet every dense subset?
h) Every dense subset of R contains a countable dense subset.
13. Arithmetic progressions. a) The arithmetic progression topology Tap (3.2)
is a topology on Z and each nonempty open set is infinite.
b) In (Z, Tap ), the basic open sets A(a, b) are also closed sets.
c) Observe that Z \ {1, −1} = {A(0, p) : p is a prime number} and deduce that
there are infinitely many primes. (This topological proof of the infinitude of primes
is due to Hillel (Harry) Furstenberg, [10].)
14. Matrices. In the space M(n, R) of all real n × n matrices,
a) the subset G L(n, R) of nonsingular matrices is open;
b) the subset S L(n, R) of matrices determinant one is closed;
c) the set Symm(n, R) of symmetric matrices is closed.
36 3 Topological Spaces
15. G δ sets and Fσ sets. A countable intersection of open sets is called a G δ -set
and a countable union of closed sets is called an Fσ -set.
a) E is a G δ if and only if E c is an Fσ ;
b) In a metric space, any closed set is a G δ and any countable set is an Fσ (In
particular, Q is an Fσ ; it is not a G δ as we will we see in the chapter on locally
compact spaces);
c) A closed interval [a, b] is both a G δ and an Fσ .
16. Regular open sets and regular closed sets. An open set V is regular open
if it is the interior of its closure: V = int V̄ and a closed set E is regular closed if it
is the closure of its interior: E = int E.
a) The closure of an open set is regular closed and the interior of a closed set is
regular open.
b) V is regular open if and only if V c is regular closed.
c) Find examples of open sets and closed sets that are not regular.
17. Kuratowski closure operation: a) Let X be a set and suppose there is a
mapping, denoted by E → cl(E), from the power set P(X ) to itself satisfying the
following properties:
K1) cl(∅) = ∅;
K2) E ⊂ cl(E);
K3) cl(cl(E)) = cl(E);
K4) cl(E ∪ F) = cl(E) ∪ cl(F) for E, F ⊂ X .
T = {E c : E ⊂ X, cl(E) = E} is a topology on X and cl(E) is just the closure Ē
in this topology. Any map satisfying the conditions K1 - K4 is called a Kuratowski
closure operation on X . [11]
b) What is an analogous operation for the interior of a set?
c) Find the Kuratowski closure operation for the cofinite topology.
d) Fix A ⊂ X and define cl(E) = E ∪ A for any nonempty E ⊂ X (and cl(∅) =
∅). This gives a closure operation. Describe the associated topology. Discuss the
cases when A = ∅, A = {a} and A = X .
18. Uniform spaces. A uniformity for a set X is a nonempty family U of subsets
of X × X with the following properties:
i) every member of U contains the diagonal = {(x, x) : x ∈ X };’
ii) if U ∈ U , then U −1 ∈ U , where U −1 = {(x, y) : (y, x) ∈ U };
iii) if U ∈ U , then V.V ⊂ U for some V ∈ U where, for V, W ∈ U , V.W =
{(x, y) : ∃z, (x, z) ∈ V, (z, y) ∈ W };
iv) if U, V ∈ U , then U ∩ V ∈ U ;
v) if U ∈ U and V ⊃ U , the V ∈ U .
A set with a uniformity is called a uniform space. This was introduced by A.Weil
[12]. For detailed treatments, see [9, 13, 14].
a) If X is a metric space, U = {U ⊂ X × X : ∃r > 0, Dr ⊂ U } is a uniformity,
where Dr = {(x, y) : d(x, y) < r }.
b) If (X, U ) is a uniform space, {E : ∀x ∈ E, ∃U ∈ U , U [x] ⊂ E} where
U [x] = {x ∈ X : (x, y) ∈ U }, is a topology on X (U -topology).
3.3 Closed Sets 37
3.4.1 Riemann
3.4.2 Weyl
Hermann Weyl (1885–1955) was a student of the great David Hilbert at Göttingen
and is well-known for his work in several areas of mathematics, including differential
operators, Riemann surfaces, Lie theory and representation theory, number theory,
foundations, relativity. He was the first to give a rigorous definition of Riemann
surfaces (in which process he used neighbourhood systems to define an abstract
space). We come across Weyl transform, Weyl character formula, Weyl integration
formula, Weyl chamber, Weyl group, Weyl basis, Peter–Weyl theorem, Schur–Weyl
duality, etc. in mathematics.
38 3 Topological Spaces
3.4.3 Hausdorff
3.4.4 F. Riesz
Frigyes Riesz (1880–1956) was a Hungarian mathematician and was one of the
founders of Functional Analysis. Besides his fundamental work in functional anal-
ysis, he contributed to integral equations, ergodic theory and topology. He made
significant contributions to axiomatic development of general topology. His main
suggestions (made in 1906 and 1908) for possible formulation were overlooked
at that time. According to historians of mathematics, his contributions came to be
appreciated only much later. For a discussion of F. Riesz’s role in the development of
general topology, see [15, 16]. The definition of connected spaces that we use today
is essentially the one given by him. Riesz Lemma, Riesz Representation Theorem,
Riesz-Fischer theorem, Riesz-Schauder theory (of compact operators), Riesz spaces,
F. and M.Riesz theorem (named after F.Riesz and his younger brother Marcel Riesz,
who is also a famous analyst) are some of the things that stand testimony to his con-
tributions to analysis. It is said that many of the central theorems of modern analysis
(e.g. Spectral theorem, Ergodic theorem) have proofs due to F. Riesz.
References
1. Riemann, B.: Über die Hypothesen, welche der Geometrie Grunde liegen. Abhandlungen der
Königlichen Gesellschaft der Wissenschaften zu Göttingen 13, 133–150 (1868)
2. Weyl, H.: Die Idee der Riemannschen Flaäche, Teubner, 1913. (Translated into English as
‘The Concept of a Riemann Surface’, (Latest: Dover 2019)
3. Hausdorff, F. Grundzüge der Mengenlehre. Leipzig (1914)
4. Tietze, H.: Beiträge zur allgemeinen Topologie I. Math. Ann. 88, 290–312 (1923). https://
doi.org/10.1007/BF01579182
References 39
Continuity is the heart of topology. Here are the first proper definitions of continuity
(on the real line).
According to a correct definition, the expression that a function f (x) varies
according to the law of continuity for all values of x inside or outside certain limits
means only that, if x is any such value the difference
f (x + ω) − f (x) can be made smaller than any given quantity, provided ω can be
taken as small as we please. (Bolzano, 1817) [1]
. . . f (x) sera fonction continue, si . . . la valeur numérique de la différence
f (x + α) − f (x) décroît indéfiniment avec celle de α . . .(Cauchy, [2] 1821).
(. . . f (x) will be (called) a continuous function, if . . . the numerical values of the
difference f (x + α) − f (x) decrease indefinitely with those of α . . .)
Wir nennen dabei eine Grösse y eine stetige Function von x, wenn man nach
Annahme einer Grösse ε die Existenz von δ beweisen kann, sodass zu jedem Wert zwis-
chen x0 − δ...x0 + δ der zugehörige Wert von y zwischen y0 − ε...y0 + ε(Weierstrass
[3, 4])
(We will call a quantity y a continuous function of x, if after choosing a quantity
ε the existence of δ can be proved, such that for any value between x0 − δ...x0 + δ
the corresponding value of y lies between y0 − ε...y0 + ε.)
Cauchy’s is an acceptable qualitative definition. The quantitative version of this
is the ε-δ definition due to Bolzano (1817) [1] and Weierstrass (1874) [4], discussed
in the chapter on metric spaces. (Some historians of mathematics feel that Cauchy
might have got the idea from Bolzano whom he met earlier in Prague during his exile
from France, see [5].)
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 41
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_4
42 4 Continuous Maps
Continuity is obviously tied to the idea of limit. We consider that later and here start
with the related concept of limit points. But first here is an example.
Example 4.1 Intuitively, the closure Ā of a set A is got by adding to A points which
are close to A. Points of A may or may not be ‘close’ to A. If A = (0, 1) ∪ {3} in
R, then Ā = [0, 1] ∪ {3}. The points 0 and 1 are not points of A but are ‘close’ to
A—all neighbourhoods of these points meet A whereas 3 ∈ A, but 3 is not close to
A—the neighbourhood (2,4) of 3 meets A in only one point, namely 3 itself.
Proof The characterisation of Ē given earlier and the definition of E make it clear
that E ⊂ Ē and so E ∪ E ⊂ Ē. On the other hand, if x ∈ Ē \ E, then, for any
neighbourhood V of x, V ∩ E = ∅, so V ∩ (E \ {x}) = ∅, and x is a limit point of
E. Now the second assertion is immediate from the fact that E is closed if and only
if E = Ē.
In metric spaces, limit points satisfy much stronger properties than the ones demanded
by the definition.
point of E, proving the first part. Hence, if x is a limit point of E, for each positive
integer n, B(x; n1 ) ∩ E is infinite. Thus xn can be chosen inductively such that xn ∈
B(x; n1 ), xn = xk , 1 ≤ k ≤ n − 1. The sequence {xn } satisfies the requirements of
the last assertion.
Corollary 4.6 A finite set in a metric space has no limit points and hence is closed.
Example 4.7 In R with the cocountable topology, any rational number q is a limit
point of the set R \ Q of irrational numbers, but no sequence of irrationals can
converge to q. So sequential limits are not sufficient to get limit points in general
topological spaces.
In spite of this ‘set back’, limit points and closures in arbitrary topological spaces
can be characterised in terms of convergence. What is needed is a concept of ‘gen-
eralised sequences’. A sequence in a set X is a map N → X where N has the natural
order. A ‘generalised sequence’ in X is got by replacing N by a set with suitable
order properties.
Definition 4.8 A directed set is a set D with an order relation ≤ which is reflexive,
transitive and has the property that any two elements have an upper bound: (i) α ≤
α; (ii) α ≤ β, β ≤ γ , α, β, γ ∈ D implies α ≤ γ ; and iii) for any pair of elements
α, β ∈ D there is a γ ∈ D with α ≤ γ , β ≤ γ . A generalised sequence or a net in a
set X is a map x : D → X from a directed set to X . Following standard convention
we write xα for x(α) and write a net x as (xα ) (just as for sequences).
Example 4.9 1. N, Q, R with their natural orders are all directed sets. Hence,
sequences are nets. (Is every E ⊂ R a directed set?)
2. The class P(X ) of all subsets of a nonempty set X is a directed set, ordered
by inclusion.
3. Let X be a topological space and let x ∈ X . The set Ux of all neighbourhoods
of x ordered by reverse inclusion (i.e. U ≤ V if V ⊂ U ) is a directed set.
4. The Cartesian product D = D1 × D2 of two directed sets, with coordinate-wise
order (α1 , α2 ) ≤ (β1 , β2 ) if α j ≤ β j , j = 1, 2.
5. Let P be the set of subdivisions S = {a = t0 < t1 < · · · < tn = b} of a closed
interval [a, b] in R. For two subdivisions S1 , S2 define S1 ≤ S2 to mean S2 is a
refinement of S1 , that is, S2 is got by adding more points of subdivisions to S1 :
S1 ⊂ S2 . This makes P a directed set.
Remark 4.11 Note that, unlike the case of sequences, ‘eventually’ may not be the
same as ‘all but finitely many’.
This convergence is also called ‘Moore–Smith convergence’, after the work (1922)
of American mathematicians Moore and Smith [6, 7]. They considered only scalar
valued ‘generalised sequences’, motivated by the example of Riemann sums, 4.12.
The concept of directed sets was independently introduced in 1921 by Vietoris [8],
who considered a general set-theoretic set up. ‘Whereas Moore and Smith con-
sidered only generalized sequences with numerical values, Vietoris studied right
from the start ... systems of sets under Zermelo’s axioms, indexed by a directed set,
and gave the definition: ... So Vietoris developed in parallel today’s theory of con-
vergence for generalized set sequences (nets) and filter bases through comparison
with the directed set of neighborhoods...Vietoris was thus the father of the modern
convergence concepts’. (Quoted with the permission of the American Mathemat-
ical Society.) [9] (For ‘filters’, another theory of convergence, see the exercises.)
Net convergence was applied to topology by the American mathematician Garret
Birkhoff (1937), [10]. The term ‘net’ is due to Kelley [11]. We will see that nets can
be used in general topological spaces just as sequences are used in metric spaces -
to characterise limit points, continuity and compactness (to be studied later).
Example 4.12 1. A net in a discrete space is convergent if and only if it is eventually
a constant. In an indiscrete space, every net converges to every point. Atrocious as
the statement sounds, it is true. Thus the indiscrete topology is anything but pleasant
or useful.
2. Take the directed set Ux of neighbourhoods of a point x in a space X and let
x V ∈ V for V ∈ Ux . The net (x V ) converges to x.
3. Let f be a bounded, real-valued function on the interval [a, b]. For any partition
P of [a, b], consider the Riemann sum s( f, P) (recall?) of f over the subdivision S.
b
The net {(s( f, P)) : P ∈ P[a,b] } converges to the Riemann integral a f (t) dt if f
is Riemann integrable. In the same way, nets of lower and upper Riemann sums can
be considered, converging to the lower and upper Riemann integrals, respectively.
Limit points and closure in metric spaces are characterised using sequences. In
topological spaces this can be done using nets.
2) Let x ∈ Ē. If x ∈ E, then the constant net (xα = x) (defined on any directed
set) converges to x. Otherwise, part (1) applies. Conversely, if x is a limit of a net
(xα ) in E, then every neighbourhood of x contains some xα and so x ∈ Ē.
The last assertion (3) follows from (2).
The basic elementary properties of the boundary are spelt out here.
4.2 Continuity
Example 4.22 Consider R with the cocountable topology Tcc and let Tseq denote
the strictly larger topology of sequentially open sets on R. So the identity map
idR : (R, Tcc ) → (R, Tseq ) is not continuous. However, a convergent sequence in
Tcc is eventually constant and so converges in Tseq as well (and, in fact, in any
topology). Thus idR preserves sequential convergence although it is not continuous.
Here are some expected results whose proofs are all trivial. Nevertheless, these
are recorded here as they are used almost all the time.
Proof (1) is trivial. Recalling the definition of the subspace topology, (2) follows as
( f | E )−1 (V ) = E ∩ f −1 (V ) for any (open) set V in Y . Lastly, (3) is a consequence
of the equality (g ◦ f )−1 (W ) = f −1 (g −1 (W )) for any (open) set W in Z .
Often one needs to ‘patch up’ together continuous maps defined on pieces, to get
a continuous map on a bigger domain. Here is one such way of gluing continuous
maps that is very useful.
Proof The identity map on any space, the inverse of a homeomorphism and the
composition of two homeomorphisms are all homeomorphisms.
unit spheres with metrics d2 , d1 and d∞ respectively. (See pictures in Chap. 2.) The
corresponding open balls Bd2 (0, 1), Bd1 (0, 1) and Bd∞ (0, 1) are also homeomorphic.
8. Consider an ellipse in R2 centred at the origin. Let A(x) denote the area enclosed
by the ellipse in the half plane {(ξ, η) ∈ R2 : ξ ≤ x}. A : R → R is a continuous
function.
9. For real numbers a, b, a < b, (a, b) = ∪∞ 1 [a + n , b). Consequently, (a, b) is
1
open in R also. So the standard topology is smaller than the lower-limit topology
on R. The identity map R → R is therefore a continuous bijection that is not a
homeomorphism.
10. Let A be an n × n real matrix. The map x → Ax on Rn given by matrix
multiplication is continuous. It is a homeomorphism if and only if A is nonsingular.
If A is orthogonal it gives a self-homeomorphism of the sphere Sn−1 , by restriction.
11. The determinant and trace are continuous functions on the matrix space
2
M(n, R) (with the Euclidean metric of Rn ) and the transpose map A → At is an
auto-homeomorphism of M(n, R).
Exercise 4
R : ξ ≤ x}. A : R → R is continuous.
2
f) Find homeomorphisms between i) the punctured plane and the exterior of the
closed unit disc; ii) the open unit disc and the open upper half plane. (Complex
numbers may make this less complex!)
g) For any space X , the set Aut(X ) of all homeomorphisms X onto itself is a
group under composition. If X and Y are homeomorphic, then Aut(X ) and Aut(Y )
are isomorphic.
h) If X = [0, 1] and Y = (0, 1), then the restriction map h → h|Y is an isomor-
phism Aut(X ) Aut(Y ).
i) Find Aut(X ) for the discrete space {0, 1} and the space S (3.2).
j) If there is a bijective continuous map from X onto Y and another from Y onto
X then, X and Y need not be homeomorphic.
k) X is homeomorphic to the diagonal (X ) = {(x, x) : x ∈ X }.
l) Suppose X, Y are countable disjoint unions of open sets, X = X n , Y = Yn .
If X n is homeomorphic to Yn for each n, then X Y .
m) Does a homeomorphism of metric spaces preserve Cauchy sequences?
n) Any homeomorphism of the sphere Sn−1 extends to a homeomorphism of the
ball Bn .
8. Isometries. a) An isometry (on a metric space) is a homeomorphism of X onto
f (X ). Find a homeomorphism that is not an isometry.
b) The set of all isometries of a metric space X onto itself is a group Iso(X ) under
composition.
c) Any isometry of R is of the form x → ax + b, a = ±1, b ∈ R.
d) On Rn , orthogonal transformations and translations are isometries. (It is a fact
that Iso(Rn ) is generated by these; this group is called the group of rigid motions of
Rn or the motion group of Rn .)
e) Let X be a set and (Y, dY ) be a metric space. Suppose σ : X → Y is a bijection.
Define dσ (x1 , x2 ) = dY (σ x1 , σ x2 ), x1 , x2 ∈ X . Then dσ is a metric on X and σ is an
isometry. If X already had a metric d X , when is dσ equivalent to d X ?
9. Semi-continuity. a) If { f i } is a family of bounded real-valued continuous
functions on a space X , then supi f i and inf i f i may not be continuous. This leads
to the concept of semi-continuous functions (due to Baire). A function f : X →
(−∞, ∞] is lower semi-continuous (lsc) if f −1 ((a, ∞]) is open in X for all a ∈ R;
f is upper semi-continuous (usc) if − f is lsc. The supremum (infimum) of any
family of lsc (usc) functions is lsc (usc).
b) If E is a subset of a space X , when is the characteristic function χ E continuous?
(χ E (x) = 1 if x ∈ E and χ E (x) = 0 if x ∈ E c .) Are there such sets in R? When is
it lower (upper) semi-continuous?
c) Every lsc function is continuous on a dense G δ [12].
10. Sequences. a) All sequences in the Sierpiǹski S (3.2) converge to 0 and the
constant sequence (1, 1, 1, · · · ) converges to both 0 and 1.
52 4 Continuous Maps
in topological spaces. (Filter bases were earlier considered by Vietoris (1931) and
Garrett Birkhoff (1935)).
A filter on a set X is a nonempty collection F of nonempty subsets of X such
that E ∩ F ∈ F whenever E, F ∈ F and F ∈ F , F ⊂ E implies E ∈ F . A filter
base E in X is a subcollection of F such that F = {F ⊂ X : ∃E ∈ E , E ⊂ F}.
(References for filters: [15, 16].)
a) Examples. i) For any x in a topological space X , the neighbourhood system
Ux at x is a filter, called the neighbourhood filter a x. A local base Bx is a filter base
for Ux . (This is the motivating example.) More generally, for any fixed E ⊂ X , the
neighbourhood filter U E at E consists of subsets A of X with E ⊂int A.
ii) For any set X and E ⊂ X , {F ⊂ X : E ⊂ F} is a filter with filter base {E}, called
the principal filter generated by E.
iii) On an infinite set X , subsets of X with finite complements form a filter. When
X = N, this is called the Fréchet filter. For n ∈ N, if Fn = {k ∈ N : k ≥ n}, then {Fn }
is a filter base for this filter.
iv) Find all filters on {1, 2, · · · , n}.
b) Convergence. We say that a filter F in a space X converges to x ∈ X , written
F → x, if Ux ⊂ F .
i) In a metric space, a convergent filter has a unique limit.
ii) In a discrete space X , F → x if and only if x ∈ F for all F ∈ F .
iii) If F is the filter formed by complements of finite sets in X , for what x does
F → x in the cofinite topology of X ?
iv) Let T1 , T2 be topologies on a set X . Then T1 is finer than T2 if and only if for any
filter F → x in T1 , F → x in T2 .
c) Closure. Let E be a subset of a space X . Then x ∈ Ē if and only if there is a
filter F with E ∈ F such that F → x.
d) Continuity and filters. If f : X → Y and F is a filter on X , let f (F ) denote
the filter with filter base all sets of the form f (F), F ∈ F. f is continuous if and
only if a filter F in converges to x ∈ X implies f (F ) → f (x) in Y .
15. Nets vs Filters. Convergence theories of net and filters are ‘equivalent’.
If (xα ) is a net in X , the associated filter is the filter generated by the filter base
consisting of sets Bβ := {xα : α ≥ β}. Conversely, if F is a filter on X , then DF =
{(x, F) : x ∈ F ∈ F } is a directed set with order (x1 , F1 ) ≤ (x2 , F2 ) if F2 ⊂ F1 .
The net ν in X defined by ν(x, F) = x is called the net based on F .
a) A net (xα ) converges to x if and only if the associated filter converges to x.
b) A filter F converges to x if and only if the net based on F converges to x.
(See, for instance, [16, 17]).
54 4 Continuous Maps
4.3.1 Cauchy
4.3.2 Weierstrass
4.3.3 Dirichlet
References
1. Bolzano, B.: Rein analytischer Beweis des Lehrsatzes, dass zwischen je zwei Werthen, die
ein entgegengesetztes Resultat gewähren, wenigstens einer reelle Wurzel der Gleichung liege.
Prague, Gottlieb Haase (1817)
2. Cauchy, A.L.: Cours d’Analyse de l’École Royale Polytechnique Paris (1821)
3. Weierstrass, K.: Über continuirliche Functionen eines reellen Arguments, die für keinen Werth
des letzteren einen bestimmten Differentialquotienten besitzen, Königl. Akad. der Wiss.,
Berlin, Werke 2, pp. 71–74 (1872)
References 57
4. Weierstrass, K.: Theorie der analytischen Funktionen, Vorlesung an der Univ. Berlin,
1874, manuscript (ausgearbeitet von G. Valentin), Math. Bibl. Humboldt Universität Berlin,
Valentin) (1874)
5. Freudenthal, H.: Did Cauchy plagiarize Bolzano? Arch. Hist. Exact Sci. 7(5), 375–392 (1971).
https://doi.org/10.1007/BF00327099
6. Moore, E.H., Smith, H.L.: A General Theory of Limits. Am. J. Math. 44, 102–121 (1922)
7. Moore, E.H.: General Analysis I. Part II. Mem American Philosphical, Society (1939)
8. Vietoris, L.: Stetige Mengen. Monatsh Math. und Physik 31, 173–204 (1921). https://doi.org/
10.1007/BF01702717
9. Reitberger, H.: Leopold Vietoris, 1891–2002. Not. Am. Math. Soc. 49(10), 1232–1236 (2002)
10. Birkhoff, G.: Moore-Smith convergence in general topology. Ann. Math. 38, 39–56 (1937).
https://doi.org/10.2307/1968508
11. Kelley, J.L.: General Topology. Affiliated East-West Press (1969)
12. Engelking, R.: General Topology Revised Edition. Heldermann Verlag (1989)
13. Hewitt, E., Ross, K.A.: Abstract Harmonic Analysis I. Springer (1963)
14. Cartan, H.: Théorie des filtres. C.R. Acad. Sci. Paris 205, 595–598 (1937)
15. Bourbaki, N.: General Topology, Chapters 1–4. Springer (1995)
16. Willard, S.: General Topology. Addison-Wesely (1970)
17. Bartle, R.G.: Nets and filters in topology. Am. Math. Monthly 62, 551–557 (1955). https://
doi.org/10.1080/00029890.1955.11988688
18. Poincaré, H.: L’oeuvre mathématique de Weierstrass. Acta Math. 22, 1–18 (1899). https://
doi.org/10.1007/BF02417867
Chapter 5
Compact Spaces
Compactness is, doubtless, the most important topological property not only for
analysis (see, e.g., [1]) but also in topology and geometry. Fréchet [2], who was the
first to define and study compactness (in metric spaces) expressed the importance of
compactness this way:
‘Nous avons déjà signalé et nous reconnaîtrons dans tout le cours de ce Livre
l’importance des ensembles compacts. Tous ceux qui ont eu à s’occuper d’Analyse
générale ont vu qu’il était impossible de s’en passer.’ (‘We have already noted, and
will recognize throughout this book, the importance of compact sets. All those who
are involved with general analysis will see that it is impossible to do without them’.)
For general topological spaces, Vietoris [3] gave the first definition thus: Eine
Menge M heiße lückenlos, wenn in ihm jede geordnete Teilmenge ohne letztes Ele-
ment mindestens einen zu M gehörigen Grenzpunkt hat. (A space is compact (lück-
enlos = ‘without gaps’) if every filter base (‘kranz’) has a cluster point, in present
terminology.)
5.1 Compactness in Rn
We begin with the classical setting of the real line, and thence to the analogous
Euclidean space setting, before moving to metric spaces and general topological
spaces. Recall the basic properties of real continuous functions f on a closed inter-
val [a, b]: (i) f is bounded and has a maximum and minimum; (ii) f is uniformly
continuous. Both these properties are of central importance. What is the underly-
ing attribute of [a, b] that leads to these properties? The fundamental characteristic
behind this nice behaviour is manifested in Borel’s remarkable result in his thesis
[4]:
Si l’on a sur une droite une infinitè d’intervalles partiels, tels que tout point de la
droite soit intérieur à l’un au moins des intervalles, on peut déterminer effectivement
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 59
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_5
60 5 Compact Spaces
un nombre limité d’intervalles choisis parmi les intervalles donnés et ayant la même
propriété- tout point de la droite est intérieur à au moins l’un d’eux. (“If on a line one
has an infinite number of subintervals, such that every point of the line is interior to
at least one of the intervals then one can determine effectively a bounded number of
intervals from among the given intervals that have the same property -every point of
the line is interior to at least one of them.” [5]) This result is remarkable, not only for
its content but also for its formulation. Its conclusion, in current terms, has become
the modern definition of compactness.
Theorem 5.1 (Borel’s theorem, 1894) Let {Vi } be a collection of open sets in [a, b]
with the property that every point of the interval is in some Vi . Then there is a finite
subcollection of {Vi } with the same property : every point of [a, b] belongs to at
least one of them. In other words, every open cover of [a, b] has a finite subcover.
Proof We present a classical proof by the method of bisection. Assume that, contrary
to the assertion of the theorem, there is an open cover {Vi } of [a, b] having no
finite subcollection that also covers I0 = [a, b]. Then this is also an open cover for
each of the intervals [a, c], [c, b], where c = a+b 2
is the midpoint of the interval. A
finite subcover for each together would give a finite subcover for the whole interval,
so for at least for one of these, call it I1 = [a1 , b1 ], there is no finite subcover.
Repeat the argument for I1 in place of I0 and choose one of the halves of I1 for
which {Vi } has no finite subcover; call it I2 = [a2 , b2 ]. Induction yields a sequence
Ik = [ak , bk ] of intervals with the following properties: Ik+1 ⊂ Ik , a ≤ ak ≤ ak+1 ≤
bk+1 ≤ bk ≤ b and bk − ak = b−a 2k
for every k. Then {ak }, {bk } are both monotonic,
bounded sequences and so have limits α = lim a k = sup ak ≤ β = lim bk = inf bk .
But bk − ak → 0 and so α = β. In other words, Ik = {α}. Now, α ∈ V j for some
j, V j is open and α = lim ak = lim bk . Thus the ak and bk eventually belong to V j ;
that is, there is a k0 such that Ik ⊂ V j for k ≥ k0 . This means a single V j covers these
Ik , contradicting the choice of the Ik . This contradiction completes the proof.
Remark 5.2 i) Here the method of bisection has been used and completeness of the
real line is the crucial ingredient, appearing in the form of Cantor’s nested interval
theorem which has been obtained in the course of the proof. This theorem says that
a decreasing sequence of closed intervals in R has nonempty intersection, and so if
the lengths of the intervals decrease to zero, then the intersection is a single point. A
proof using the equivalent lub, glb properties can be given (as was done in the proof
of the intermediate value theorem; bisection method can be used to prove that result
too.)
ii) In the theorem, it is immaterial whether the Vi are assumed to be open in [a, b]
or in R (Why?).
Definition 5.3 A topological space X is said to be compact if every collection of
open sets covering it has a finite subcollection which also covers it; i.e. every open
cover for X has a finite subcover.
5.1 Compactness in Rn 61
The history of compactness and related properties in the real line is complicated and
we will mention only a few salient points. (For more details, see Manya Raman-
Sundström [5] and Andre et al. [6].) Borel stated his result 5.1 only for countable
covers and the general form is usually credited to the famous French mathematician
Henri Lebesgue, who gave a short elegant proof of the general result. Almost at the
same time as Borel, Pierre Cousin (of France, 1867–1933) had also proved (1895)
the general form of Borel’s theorem (5.1). Perhaps the result may be called the
Borel–Cousin–Lebesgue theorem, but there are several other names involved; see
the references mentioned above.
The proof of compactness
of intervals [a, b] easily extends to give compactness
of ‘rectangular boxes’ n1 [a j , b j ], in Rn .
Theorem 5.4 Any box n1 [a j , b j ] in Rn is compact.
The stage is now set for presenting the famous characterisation of compact subsets
of Rn . It is essentially a consequence of Borel’s theorem. Heine’s name had been
mistakenly associated with the theorem. To quote from [5]: “While Heine is credited
with a theorem he did not prove, it appears that Cousin was largely overlooked for
a lemma he did prove.” (© Mathematical Association of America, 2015. All rights
reserved. Quoted here with permission.) But the name Heine–Borel theorem has now
become entrenched, though many French texts call it Borel–Lebesgue theorem. As
a still incorrect compromise, we combine all the three names.
Proof Boundedness is with respect to the Euclidean metric d2 (or, equivalently, the
‘box’ metric d∞ ). Thus a set K in Rn is bounded if and only if it is contained in some
Euclidean ball or in some box.
If K is not bounded, there are points of K outside any ball and so the open cover
{B(0; k) : k = 1, 2, ...} does not have a finite subcover. Thus K is not compact. On
the other hand, if K is not closed, there is a limit point x0 of K with x0 ∈ / K . Then
every ball B(x0 ; k1 ), k = 1, 2, ... contains points of K and the complements of the
closed balls B̄(x0 ; k1 ) form an open cover for K without any finite subcover. Thus
again K is not compact.
Conversely, suppose K is closed and bounded. It is then a closed subset of some
large enough box B. If {Vi } is an open cover for K , then the Vi together with the
62 5 Compact Spaces
open set K c is an open cover for the set B which is compact by the earlier result.
A finite subcover obtained from this compactness also covers K and the theorem is
proved.
Example 5.6 1. R, (0, 1), (a, ∞), (−∞, b] are not compact.
2. Closed balls and spheres in Rn are compact; half planes in R2 and hyperboloids
in R3 are not.
3. The Cantor ‘middle-third’ (or ‘ternary’) set in R is compact.
4. The set {(a,
b, c, d) ∈ R4 : ad − bc = 1} is noncompact, whereas
a b
{(a, b, c, d) ∈ R4 : is an orthogonal matrix} is compact.
cd
5. The previous example easily generalises. The space S L(n, R) of real n × n
matrices of unit determinant is noncompact and the set O(n) of orthogonal matrices
is closed and bounded, hence is compact.
Here is the first ever definition of compactness, although it was given in the context
of metric spaces. Fréchet’s definition gives what we now call ‘limit point compact
spaces’.
Nous dirons qu’un ensemble est compact lorsqu’il ne comprend qu’un nombre
fini d’éléments ou lorsque toute infinité de ses éléments donne lieu à au moins un
élément limite. (Fréchet, [7])
(‘We say that a set is compact if it is finite or any infinitely many of its elements
has at least one limit point.’)
The open cover definition of compactness is perhaps a bit mysterious and is
hardly appealing to intuition. In the case of Rn , the Heine–Borel–Lebesgue theorem
gives a more tangible characterisation of compact sets. In the case of metric spaces,
Fréchet’s formulation and other, intuition friendly and simpler to apply, formulations
are equivalent to the open cover definition.
Recall: (i) Every bounded sequence of real numbers has a convergent subsequence;
(ii) every bounded infinite set of real numbers has a limit point. If a subset K of R
is compact, then it is bounded and closed and so (1) every sequence in K has a
subsequence convergent in K and (2) every infinite subset of K has a limit point in
K . It turns out that each of these properties characterises compactness in arbitrary
metric spaces, not just in R. Thus we are led to the following convenient definitions.
Bolzano’s result on the greatest lower bound property enabled Weierstrass to prove
that infinite bounded sets of real numbers have limit points (Bolzano–Weierstrass
theorem). Fréchet used this property to define limit point compactness for metric
spaces.
Our first objective is to prove the equivalence of each of these with compactness in
metric spaces. It is not difficult to get the mutual equivalence of these two properties
and the implication that these properties are implied by compactness. The hard part is
to prove that sequentially compact metric spaces are compact. In the process we also
encounter another important and interesting property for subsets of metric spaces that
is stronger than boundedness. We take this up now, beginning with some examples.
Example 5.8 1. Consider an interval (a, b), a, b ∈ R, a < b. It is of course bounded,
but there is something more. To see this, let ε > 0 and choose a positive integer n
such that b−an
< ε and subdivide the interval into n equal subintervals. Then each
subinterval is of length less than ε. Another way of stating this is that there are finitely
many points (namely, the points of subdivision) such that every point of the whole
interval is at a distance less than ε from at least one of these points. (It is immaterial
whether the interval is open, closed, half-open or half-closed for this discussion.)
2. On the other hand, consider an infinite set with discrete metric. The distance
between two distinct points is always one and the phenomenon described in the
previous example is far from true here.
3. For a nontrivial space of the kind encountered in the previous example, consider
the space ∞ of all bounded (real or complex) sequences with the metric d(x, y) =
sup |xn − yn |, x = (xn ), y = (yn ) ∈ ∞ . Let
E = {e1 = (1, 0, 0, ...), e2 = (0, 1, 0.0, ...), e3 = (0, 0, 1, 0, ...), ...}.
Then d(en , em ) = 1, n = m. E is bounded, but the property described in Example 1
fails miserably here too. What happens in 1 and 2 ?
This discussion leads us to a strong form of boundedness that is related to com-
pactness and is especially useful in analysis.
Definition 5.9 A subset E of a metric space X is said to be totally bounded if, for
every ε > 0, there is a finite set of points x1 , ..., xk in X such that for any x ∈ E
there is an x j with d(x, x j ) < ε. This just means that finitely many balls of any given
radius cover E. The set of centres of such a collection of balls, {x j }k1 , is called an
ε-mesh for E.
The examples above show that any bounded interval (and hence any bounded set)
in R is totally bounded (what happens in Rn ?), whereas an infinite set in a discrete
metric space and the set of unit vectors in ∞ are bounded, but far from totally
bounded. In fact, we have seen that for any metric d on a set X , there is an associated
equivalent bounded metric d ∗ .
For any ε > 0, a metric space can be covered by ε-balls, and if the space is compact,
a finite subcollection of these balls cover the space. Thus there appears to be some
resemblance between compactness and total boundedness. However, [a, b], (a, b)
64 5 Compact Spaces
are both totally bounded, whereas only the former is compact. What makes the
difference? An obvious difference between the two intervals is that [a, b] is closed in
R and so is complete but (a, b) is not. This turns out to be the crucial difference. In
fact, total boundedness and completeness, in tandem, yield compactness. To obtain
this last mentioned implication, we shall invoke an interesting covering result, also
called Lebesgue Number Lemma, that is of independent importance.
Theorem 5.10 (Lebesgue covering lemma) Let {Vi } be an open cover of a sequen-
tially compact metric space X . There is a δ > 0 (called a Lebesgue number of the
cover {Vi }) such that every subset E of X with diameter at most δ is contained in
some Vi .
Proof Suppose, on the contrary, that there is no such δ. This means that for every
δ > 0 there is a set E with diam E < δ which is not contained in any Vi . Take δ = n1
and denote the corresponding E by E n , n = 1, 2, .... Choose a sequence (xn ) such
that xn ∈ E n for every n. By the assumed sequential compactness of X , there is a
subsequence (xn k ) which converges to some x ∈ X . Since {Vi } is a cover for X , there
is a j with x ∈ V j and so V j contains a ball B(x; ε). Choose k large enough so that
1
nk
< 2ε and xn k ∈ B(x; 2ε ). Then, for y ∈ E n k we have
ε ε
d(y, x) ≤ d(y, xn k ) + d(xn k , x) < 2
+ 2
= ε.
Thus E n k ⊂ B(x; ε) ⊂ V j , contrary to our assumption on the E n . So our assumption
is untenable and the proof is complete.
The definition of compactness that each open cover has a finite subcover can be recast
in terms of closed sets, by taking complements: every collection of closed sets having
empty intersection has some finite subcollection also having empty intersection. This
will be very useful.
A sequence {xn } in a metric space (X, d) is called a Cauchy sequence (or a d-
Cauchy sequence) if for every ε > 0 there is an n ε > 0 such that d(xm , xn ) < ε for
m, n > n ε . X is said to be complete (or d-complete) if every Cauchy sequence in X
has a limit in X . Completeness is a very important property and a later chapter is fully
devoted to it. For now, we use it in the following immensely useful characterisation.
Theorem 5.11 For a metric space X , the following are equivalent:
1) X is compact;
2) X is Bolzano–Weierstrass compact;
3) X is sequentially compact;
4) X is totally bounded and complete.
Proof 1) ⇒ 2). Suppose there is an infinite set E in X having no limit point. Then
for each x ∈ E, E x := E \ {x} also has no limit point and so is closed. Every finite
subcollection of the collection {E x : x ∈ E} of closed sets has nonempty intersection
as E is infinite. But, clearly, ∩x∈E E x = ∅. The conclusion is that X is not compact.
2) ⇒ 3). Assume that every infinite subset of X has a limit point, and let {xn }
be a sequence in X . If the sequence has only finitely many distinct points, then it
5.2 Compactness in Metric Spaces 65
has a subsequence that is constant and hence convergent. Otherwise the set of points
of the sequence is infinite and so has a limit point x, by hypothesis. So, there is an
xn k ∈ B(x; k1 ) for each k with n 1 < n 2 < · · · . Then {xn k } converges to x.
3) ⇒ 4). If a Cauchy sequence has a convergent subsequence, then the original
sequence itself converges. This shows that sequential compactness of X immediately
implies completeness. We assume X is not totally bounded, so there is an ε > 0 such
that X has no finite ε-mesh. We prove that it is not sequentially compact. For any
x1 ∈ X there is an x2 with d(x1 , x2 ) ≥ ε. By induction, there is a sequence {xn }
satisfying the lower bound d(xm , xn ) ≥ ε whenever m = n. Clearly no subsequence
of {xn } can be convergent.
4) ⇒ 3). Assume that X is totally bounded and complete. To prove sequential com-
pactness, it suffices to show that every sequence {xn } in X has a Cauchy subsequence.
Cover X by a finite number of balls of radius 1. One of these balls, B1 , contains xn
for infinitely many n, so contains a subsequence, denoted by {x1n }. Repeating the
argument for balls of radius 21 with the sequence {x1n } in place of {xn }, we conclude
that there is a ball B2 of radius 21 containing a subsequence {x2n } of {x1n }. Inductively,
for each positive integer k > 1 we get a ball Bk of radius k1 containing a subsequence
{xkn } of {x(k−1)n } (so d(xkn , xkm ) < 2/k for all n, m). Then the ‘diagonal’ subse-
quence {xnn } is a Cauchy sequence. (Why?) (The method employed here exemplifies
what is usually referred to as ‘Cantor’s diagonal argument’. It is a simple, yet pow-
erful tool that is useful in diverse situations and the reader will do well to master it.)
3) + 4) ⇒ 1). Here is where we need to invoke the Lebesgue covering lemma.
Since 3) and 4) are equivalent, this implication would complete the proof. Suppose
X is sequentially compact, hence is totally bounded. Let {Vi } be an open cover for
X and let δ > 0 be a Lebesgue number for this cover. By total boundedness, a finite
number of balls B(x j ; 2δ ), 1 ≤ j ≤ k, cover X . Since the diameters of these balls
are less than δ, for each j there is an i j with B(x j ; 2δ ) ⊂ Vi j . Hence Vi j , 1 ≤ j ≤ k,
form a cover for X . The proof is complete.
The equivalence (for metric spaces) of sequential compactness with limit point com-
pactness and with the open cover definition were obtained by Fréchet and Hausdorff,
respectively.
Remark 5.12 Using this characterisation, another proof of the Heine–Borel–
Lebesgue theorem can be given: Any bounded sequence of real numbers has a conver-
gent subsequence. So, if K is a closed and bounded set in Rn , any sequence from K
will give n bounded sequences of real numbers, one for each coordinate, and each of
these will have convergent subsequences. Convergence in Rn is coordinate-wise con-
vergence and we can get a convergent subsequence of the original sequence. Thus K
is sequentially compact. Alternately, the Bolzano–Weierstrass theorem for real num-
bers implies that any closed and bounded set in Rn is Bolzano–Weierstrass compact.
An important property of a continuous function on [a, b] is that it is uniformly
continuous. Dirichlet had observed this in his lectures in 1852. Heine ([8], 1870)
mentions uniform continuity as follows.
66 5 Compact Spaces
Es scheint aber noch nicht bemerkt zu sein, dass . . . diese Continuität in jedem
einzelnen Punkte . . . nicht diejenige Continuität ist . . . die man gleichmässige Conti-
nuität nennen kann, weil sie sich gleichmässig Über alle Punkte und alle Richtungen
erstreckt.
(‘It does not seem to have been observed yet, that . . . continuity at each single
point . . . is not that continuity . . . which can be called uniform continuity, because
it extends uniformly to all points and in all directions.’
We give a brief ab initio discussion on uniform continuity here. We look at a
couple of examples first, to motivate and clarify the concept.
Example 5.13 1. Recall the ε-δ definition of continuity. Consider the following
function: f : R → R, f (x) = sin x. Observe that
| sin x − sin y| = |2 cos( x+y
2
) sin( x−y
2
)| ≤ 2| sin( x−y
2
)| ≤ 2 |x−y|
2
= |x − y|
if |x − y| is small enough. Hence, | sin x − sin y| < ε for any sufficiently small ε,
for any two real numbers x, y with |x − y| < δ, where δ is any positive real number
with δ ≤ ε. In other words, we get, for a given ε, a δ independent of the points being
considered.
2. On the other hand, consider the function f (x) = x1 on (0,1). Let us try to
prove its continuity. Take x0 ∈ (0, 1) and let ε > 0. We want to find a δ > 0 such
that | f (x) − f (x0 )| < ε whenever |x − x0 | < δ. To begin with, choose 0 < δ <
x0
2
. If |x − x0 | < δ, then x20 < x < 3x20 and so x1 < x20 . Thus | x1 − x10 | = |x−x
x x0
0|
<
x02
2δ
x02
. The last term is smaller than ε provided δ < 2
ε. Hence |x − x0 | < δ implies
x02
| f (x) − f (x0 )| < ε if 0 < δ ≤ min{ x20 , 2 ε}. The important thing to observe is that
this δ gets smaller and smaller as x0 comes closer and closer to 0. In fact, it is not
possible to find a δ which is independent of x0 . For instance, for any 0 < δ < 1, we
δ δ2 δ
have 0 < δ − 1+δ = 1+δ < δ whereas | f (δ) − f ( 1+δ )| = | 1δ − 1+δ
δ
| = 1. Another
way is to observe that n − 2n = 2n is small if n is large whereas | f ( n ) − f ( 2n
1 1 1 1 1
)| = n
becomes unbounded, so a choice of a uniform δ is impossible.
Note, however, if η is any fixed positive real number, then
| x1 − x10 | = |x−x
x x0
0|
< |x−x
η2
0|
for x, x0 ∈ (η, 1),
so that if δ > 0 is chosen with δ ≤ η2 ε, then | f (x) − f (x0 )| < ε whenever
|x − x0 | < δ whatever x0 ∈ (η, 1) is.
The point of these examples is to show that δ depends, in general, not just on ε,
but also on the point x0 being considered. When it is possible to choose a uniform δ
for all points of the space, depending only on ε, we have a strong form of continuity.
reason for the uniform continuity of functions on [a, b] is its compactness. Now we
can justify this statement.
Uniform continuity of continuous functions on [a, b] was first proved by Dirichlet
‘(with a more explicit use of coverings and subcoverings than in Heine’s theorem’
[5] (© Mathematical Association of America, 2015. All rights reserved. Quoted here
with permission.)) as early as 1852 in his lectures, which appeared in print only in
1904. In the meanwhile Heine, who studied under Dirichlet, published a proof in
1872 [9].
Theorem 5.15 (Uniform continuity theorem) Let (X, d) be a compact metric space
and (X
, d
) be an arbitrary metric space. Then any continuous map f : X → X
is
uniformly continuous.
Proof 1. This is an easy consequence of the Lebesgue covering lemma. Let ε > 0
and cover X
by open balls Bi of radius 2ε . Then the pull-backs f −1 (Bi ) form an open
cover for X . If δ is a Lebesgue number for this cover, then x, y ∈ X , d(x, y) < δ
imply x, y ∈ f −1 (B j ) for some j. Hence f (x), f (y) ∈ B j , so d
( f (x), f (y)) ≤
diam B j < ε.
Proof 2. Assume f : X → X
is continuous, but not uniformly. Then there is an ε >
0 such that for every positive integer n there are points xn , yn ∈ X with d(xn , yn ) <
1
n
, but d
( f (xn ), f (yn )) ≥ ε. By sequential compactness, the sequences {xn } and
{yn } have convergent subsequences {xn k } and {yn k }, say xn k → x, yn k → y in X .
(Why can we choose a common subsequence {n k }?) But d(xn k , yn k ) → 0, hence
x = y. Then f (xn k ) → f (x) and f (yn k ) → f (x) in X
and so d
( f (xn k ), f (yn k )) →
d( f (x), f (x)) = 0. This is impossible because d
( f (xn k ), f (yn k ) ) ≥ ε for all k. The
proof is complete.
Corollary 5.16 If X is a compact metric space (e.g. X = [a, b]) and f : X → R is
continuous, then f is uniformly continuous.
The second proof provides us with a useful sequential criterion.
Proposition 5.17 Let f : (X, d) → (X
, d
) be a map from one metric space to
another. The following conditions are equivalent:
i) f is uniformly continuous;
ii) d
( f (xn ), f (yn )) → 0 whenever d(xn , yn ) → 0.
Proof Let f be uniformly continuous and d(xn , yn ) → 0 where xn , yn ∈ X . Let
ε > 0 and choose a corresponding δ from the uniform continuity of f . Choose n δ
such that n ≥ n δ implies d(xn , yn ) < δ. Thus n ≥ n δ implies d(xn , yn ) < δ implies
d
( f (xn ), f (yn )) < ε.
If (i) fails, then for some ε > 0 and for every n, there are xn , yn ∈ X with
d(xn , yn ) < n1 and d
( f (xn ), f (yn )) ≥ ε, so (ii) fails.
Exercise. Prove that a uniformly continuous map takes Cauchy sequences to
Cauchy sequences. Is the converse true? (Hint: If {xn } is a Cauchy sequence in
R, what about {xn2 }?) What if f is continuous?
Exercise. Use the previous exercise to show that the function 1/x on (0,1) is not
uniformly continuous.
68 5 Compact Spaces
Although the first compactness result of Borel [10, 11] (5.1) was formulated in terms
of open covers in the 1880’s, the general definition was proposed by Tietze [12] and
adopted by Alexandroff and Urysohn [13, 14] (who used the term ‘bicompact’) only
much later, in the 1920’s. This modern open cover definition appeared in Alexan-
droff and Hopf [15]. Ein topologischer Raum heißt bikompakt, wenn jede seiner
offenen Überdeckungen eine endliche Überdeckung enthält. (‘A topological space is
bicompact if each of its open cover contains a finite cover.’)
Proof Let {Vi } be any open cover for X . Writing each Vi as a union of members of
B, we get a cover of X by members of B. By hypothesis, this has a finite subcover
{B1 , ..., Bk } of basic open sets. Choosing, for each j, 1 ≤ j ≤ k, a Vi j such that
B j ⊂ Vi j we end up with a finite subcover {Vi1 , ..., Vik } of the original cover {Vi }.
Here is a definition that will enable us to state the already mentioned reformulation
of compactness in terms of closed sets.
Proposition 5.21 (F. Riesz) A topological space is compact if and only if every
family of closed subsets having finite intersection property has nonempty intersection.
Proof As already observed, just take complements in the open cover definition.
Write out the details.
Exercise. Prove the corollary from the open cover definition directly.
Exercise. Prove the following variant of the proposition: X is compact if and only
if Āi is nonempty for every family {Ai } of subsets of X with finite intersection
property.
Exercise. A subset K of a space X is compact if and only if every collection of
sets open in X covering K has a finite subcover. Thus it is immaterial whether sets
of the cover are open in K or in X .
Here is a simple, widely used property of compact spaces.
Proof If {Vi } is an open cover for f (X ), then the pull-backs f −1 (Vi ) form an open
cover for the compact space X and so finitely many of these { f −1 (Vi j )}k1 cover X .
Then the Vi j cover f (X ).
The classical theorem on maxima and minima of real continuous functions on closed
intervals is now an easy consequence.
The case X = [a, b] is a basic result in Real Analysis and was first published by
Weierstrass in 1877, although he must have proved and used it in his lectures much
earlier. The ideas underlying his proof go back to the proof of a fundamental property
of real numbers, namely, the greatest lower bound property, which Bolzano published
in 1817.
Corollary 5.25 If f is a continuous bijection from a compact space X onto a metric
space Y , it is actually a homeomorphism.
Proof To show that f −1 , which exists as f is bijective, pulls back closed sets to
closed sets, i.e. to show that f maps closed sets to closed sets. If C is closed in X ,
then it is compact and so f (X ) is compact. As Y is a metric space, the compact set
f (C) is closed.
The proof hinges on the fact that compact sets are closed in metric spaces. So, the
result holds if the range space has this property. Hausdorff spaces are such more
general spaces. See 8.10.
Recall that a metric space X is compact if and only if every sequence {xn } has a
convergent subsequence {xn k }. The statement that {xn } has a subsequence xn k → x
can be reformulated as follows: the xn are frequently in every neighbourhood U of
70 5 Compact Spaces
x, i.e. for any N there is an n > N such that xn ∈ U . In this case we say that x is a
cluster point of the sequence {xn }. So x is a cluster point of {xn } if and only if there
is a subsequence converging to x. Hence a metric space X is compact if and only if
every sequence in X has a cluster point.
This is no longer true for a general topological space. But is there an analogous
result for arbitrary topological spaces? Indeed, we have an exact, useful analogue
in terms of nets. The concept of a cluster point of a net poses no problems. But
what is the analogue of a subsequence? A subsequence of the sequence n → xn
is got by composing this map (i.e. the sequence) with a strictly increasing map
k → n k , N → N, to get the sequence k → xn k . This leads us to subnets. One last
point to be remembered - the domains of definition of a sequence and a subsequence
are always the same set N. But for a net and a subnet they can be different directed
sets.
Definition 5.26 A point x ∈ X is a cluster point of a net {xα } in X if xα lies frequently
in every neighbourhood Ux of x. This means {α : xα ∈ Ux } is a cofinal subset. Here,
a subset E of a directed set D is said to be cofinal if for every α ∈ D there is a
β ∈ E with β ≥ α. A subnet of a net x : D → X is a net x ◦ ν : E → X , where E is
a directed set and ν : E → D is a map which satisfies the condition: for each β ∈ D
there is an η ∈ E such that ξ ≥ η implies ν(ξ ) ≥ β.
In particular, if ν is increasing (ν(η) ≤ ν(ξ ) if η ≤ ξ ) and ν has cofinal range, then
x ◦ ν is a subnet. In consonance with our notation for subsequences, we write {xαη } for
the subnet that is the composition of the net α → xα and the map η → αη , E → D.
The condition on ν says that ν(ξ ) is ‘large’ if ξ is ‘large’. Kelley [17] coined the term
net and introduced subnets in 1950. Sequences are nets, subsequences are subnets
and the cofinality condition is automatic in the case of sequences. (Caution: a subnet
of a sequence need not be a subsequence.)
As early as 1921, Vietoris [3] introduced the concept of a cluster point (‘Gren-
zpunkt’): “Definition: Ein Punkt r heisst teilweiser Grenzpunkt einer orientierten
Menge zweiter Ordnung M ... ohne letztes Element wenn es zu jeder Umgebung
V von r und zu jedem Element Q β von M noch Elemente Q α > Q β gibt, welche
Punkte von V enthalten.” (A point r is a cluster point of M if for any neighbourhood
V of r and any Q β ∈ M, V contains points Q α > Q β .)
Proposition 5.28 A point x in a space X is a cluster point of a net {xα } if and only
if it is a limit of some subnet of {xα }.
Proof If x is not a cluster point of {xα }, then there is a neighbourhood U of x such that
{xα } is not frequently in U , so is eventually in U c . Then any subnet is also eventually
5.3 Compactness in Topological Spaces 71
ordinal space. Since we have not discussed ordinal numbers, we refer to our standard
references [18, 19, 21, 22] for examples involving ordinal spaces.
11. Noncompact balls. The closed unit balls in the sequence spaces 1 , 2 and
∞
and in C[0, 1] are not compact.
12. Semi-continuity. If f is a real lower (upper) semi-continuous on a compact
space, then f has a minimum (maximum) on X .
13. Bolzano–Weierstrass. a) On N, consider the topology generated by {Un =
{2n − 1, 2n}}n∈N as a subbase. This space is Bolzano–Weierstrass compact, but not
compact nor sequentially compact.
b) Does the Bolzano–Weierstrass theorem hold in R ?
14. Compact set of eigenvalues. The set of eigenvalues of a compact set K of
matrices is compact: i.e. if σ (A) is the set of eigenvalues of an A ∈ M(n, R), then
σ (K) := ∪{σ (A) : A ∈ K} is compact.
15. Ultranets. A net {xα } in X is called an ultranet or a universal net if for any
E ⊂ X , {xα } is either eventually in E or eventually in E c . See Kelley [18] or Willard
[19] for this exercise and the next.
a) The image of an ultranet under any map is an ultranet.
b) Every net has a subnet which is an ultranet.
c) A subnet of an ultranet is an ultranet.
d) A space X is compact if and only if each ultranet in X converges.
16. Ultrafilters. A filter F on X is called an ultrafilter if it is a maximal filter,
i.e. there is no filter F ∗ with F F ∗ . [23]
a) A filter F on X is an ultrafilter if and only if for any E ⊂ X, either E ∈ F or
Ec ∈ F .
b) Every filter is contained in an ultrafilter. (Needs Zorn’s lemma.)
c) Any map takes an ultrafilter to an ultrafilter.
d) X is compact if and only if every ultrafilter in X converges if and only if every
filter in X has a cluster point. (A point x is a cluster point of a filter F if each F ∈ F
intersects every U ∈ Ux .)
e) Obtain results on the relation between cluster points of nets and filters and
between ultranets and ultrafilters. (see e.g. [19])
5.4.1 E. Borel
Émil Borel (1871–1956) was a French mathematician and politician. His main math-
ematical work was in analysis, measure theory and probability. Borel’s theorem on
the compactness of closed intervals, Heine–Borel–Lebesgue theorem, Borel sum-
mation, Borel sets, Borel measure, Borel algebra, Borel–Cantelli lemma, Borel’s
Law of Large Numbers, Borel–Carathéodory theorem, Borel distribution bear his
74 5 Compact Spaces
name. The famous mathematician Lebesgue was his student. There is the Centre
Émil Borel at the Institut Henri Poincaré (the institute of which he was a founder).
A moon crater is named after him. He was a cabinet minister for a brief period when
the mathematician Paul Painlevé was the Prime Minister.
5.4.2 Heine
Eduard Heine (1821–1881) was a German mathematician known for his work on
Real Analysis and special functions. He dedicated his thesis (1842), on differential
equations, to Dirichlet. His lectures were known for clarity and as a teacher he was
greatly liked. He was the first to publish a proof of uniform continuity of continuous
functions on closed intervals. He is famous because of the Heine–Borel theorem.
5.4.3 Lebesgue
5.4.4 Cantor
Georg Cantor (1845–1918) was the creator of Set theory on which the structure
of modern mathematics has been raised. His theories of ‘infinity of infinities’ and
transfinite numbers and their arithmetic were vehemently opposed by many leading
mathematicians of his time, including his teacher Leopold Kronecker. This hos-
5.4 Biographical Notes 75
tility greatly affected his work and is thought to be the reason for several bouts of
mental depressions he suffered from 1884 until his death. This criticism was matched
by later encomium (e.g. David Hilbert’s famous quotes: ‘No one shall expel us from
the paradise that Cantor has created.’ Cantor’s set theory is ‘The finest product
of mathematical genius and one of the supreme achievements of purely intellectual
human activity.’ https://mathshistory.st-andrews.ac.uk/Biographies/Cantor/). The
Royal Society’s highest honour was bestowed on him. Initially, he wrote several
papers on number theory, before he turned to Trigonometric Series. He solved the
long-standing problem of the uniqueness of trigonometric series representation of
functions and, in the course of this work, discovered transfinite ordinals. Cantor’s
1874 paper on algebraic numbers marked the beginning of set theory. He showed that
algebraic numbers are countable, real numbers are not and so deduced the existence
of (uncountably many) transcendental numbers, without producing even one tran-
scendental number. This kind of proofs were also criticised. He proved the existence
of many kinds of infinities, developed the theory of transfinite numbers, cardinal
numbers, ordinal numbers, well-ordering principle, introduced the concept of one–
one correspondence (His proof that ‘the unit interval contains as many points as
the unit square’ shocked the mathematical world.), the diagonal argument (which
is simple, powerful and very useful; we have used it.), got the real numbers as the
completion of rationals (before Richard Dedekind developed his ‘cuts’ to get real
numbers), formulated the Continuum Hypothesis, which he tried to prove in vain
and was the first of the famous 23 problems that Hilbert listed, at the Paris Inter-
national Congress (1900); it has since been shown that it can neither be proved nor
disproved in the standard ZFC-set theory, as Kurt Gödel (1940) and the harmonic
analyst turned logician Paul Cohen (1963) showed. Kronecker and Weierstrass
were his thesis advisers. Cantor set, Cantor intersection theorem, Cantor function
are well-known in analysis and topology. Cantor was instrumental in the formation
of the German Mathematical Society and also in convening the first International
Congress of Mathematicians.
5.4.5 Vietoris
Leopold Vietoris (1891–2002) (Look at the years! No typo!) was an Austrian mathe-
matician, mainly known for his contributions to topology, both general and algebraic,
(and for his longevity! His last paper appeared when he was 103!). He obtained his
first results when he was fighting in World War I, was an Italian prisoner of war
and, soon after his release, submitted his thesis to the University of Vienna in 1919.
He spent three semesters in Amsterdam with Brouwer; Alexandroff and Menger
were also there at that time. During his stay there, he obtained his most significant
results in Algebraic Topology—Mayer–Vietoris sequence for homology groups, etc.
He defined regularity of topological spaces (slightly different from the currently
used definition due to Tietze) and the modern notion of compact spaces (with the
76 5 Compact Spaces
name ‘spaces without gaps’). As mentioned earlier (after Remark 4.11), he came up
with the idea of ‘directed sets’ in a general setting (1921) even before the Moore-
Smith paper on nets (1922) and also the concept of a filter base. He had essentially
developed a theory of convergence parallel to the modern theories of net and filter
convergence and so may be considered as the originator of modern convergence
theory (see Heinrich Reitberger [24]). He essentially proved (1921) the normality
of compact Hausdorff spaces. He made attempts to formulate the definition of a
manifold. The ‘Vietoris topology’ that he defined on the class of closed sets in a
topological space coincides with the Hausdorff metric on compact metric spaces.
Mayer–Vietoris sequences, Vietoris complexes, Vietoris cycles, Vietoris mapping
theorem are some other concepts in algebraic topology. Functional differential equa-
tions, probability theory, orientation in mountainous terrain by differential geometric
methods, the strength of the alpine ski and the physics of block glaciers are other
areas to which he made contributions.
References
1. Hewitt, E.: The role of compactness in analysis. Am. Math. Monthly 67, 499–516 (1960).
https://doi.org/10.2307/2309166
2. Fréchet, M.: Espaces abstraits. Gauthier-Villars, Paris (1928)
3. Vietoris, L.: Stetige Mengen. Monatsh Math. und Physik 31, 173–204 (1921). https://doi.org/
10.1007/BF01702717
4. Borsuk, K.: Sur les rétractes. Fund. Math. 17, 152–170 (1931). https://doi.org/10.4064/fm-17-
1-152-170
5. Raman-Sundström, M.: A pedagogical history of compactness. Am. Math. Monthly 122, 619–
635 (2015). https://doi.org/10.4169/amer.math.monthly.122.7.619
6. Andre, N., Engdahl, S., Parker, A.: An analysis of the first proofs of the Heine-Borel Theorem-
Borel’s Proof”. Loci (2003). https://doi.org/10.4169/loci003890
7. Fréchet, M.: Sur quelques point du calcul fonctionnel. Rendiconti di Palermo 22, 1–74 (1906).
https://doi.org/10.1007/BF03018603
8. Heine, E.: Über trigonometrische Reihen. J. Reine und Angew. Math. 71, 353–365 (1870).
https://doi.org/10.1515/crll.1870.71.353
9. Heine, E.: Die Elemente der Funktionenlehre. J. Reine und Angew. Math. 74, 172–188 (1872).
https://doi.org/10.1515/crll.1872.74.172
10. Borel, E.: Sur quelques points de la théorie des fonctions. Gauthier-Villars (1894). https://doi.
org/10.24033/asens.406
11. Borel, E.: Leçons sur la Théorie des Fonctions. Paris (1898)
12. Tietze, H.: Beiträge zur allgemeinen Topologie I. Math. Ann. 88, 290–312 (1923). https://doi.
org/10.1007/BF01579182
13. Alexandroff, P., Urysohn, P.: Zur theorie der topologischen Räume. Math. Ann. 92, 258–266
(1924). https://doi.org/10.1007/BF01448008
14. Alexandroff, P., Urysohn, P.: Memoire sur les espaces topologiques compactes. Verh. Akad.
Wetensch. Amsterdam 14, 1–96 (1929)
15. Alexandroff, P., Hopf.: Springer, H. Topologie I (1935)
16. Bourbaki, N.: General Topology, Chapters 1–4. Springer (1995)
17. Kelley, J.L.: Convergence in topology. Duke Math. J. 17, 277–283 (1950). https://doi.org/10.
1215/S0012-7094-50-01726-1
18. Kelley, J.L.: General Topology. Affiliated East-West Press (1969)
References 77
The product topology is a very important special case of an important way of defining
topologies using maps: defining a topology on a set X using maps from X to topo-
logical spaces. In the other direction, getting a topology on a set Y from maps into Y
from spaces has quotient topology as a particular case. We discuss both these ways of
defining topologies, (called the initial topology and the final topology, respectively,
by Bourbaki) before specialising to the product and quotient topologies.
Tychonoff’s famous 1930 definition of the product topology is given only for
arbitrary products of copies of [0,1]:
Es sei {Jα } eine Menge von der Mächtigkeit τ von abstrakt gegebenen zueinan-
der fxemden Einheitsstrecken 0 ≤ t ≤ 1. Ein Punkt x des Raumes Rτ isr definition-
sgemäss der Inbegriff {t1 , t2 , . . . , tα , . . .} von “Koordinaten” tα , wobeit tα ein Punkt
vonJα , also eine reelle Zahl 0 ≤ tα ≤ 1 ist. Die Umgebungsdefinition geschieht fol-
gendermassen: Es sei x0 = {t10 , t20 , . . . , tα0 , . . .} ein Punkt yon Rτ . Wir wählen beliebig
endlichviele Jα1 , Jα2 , . . . , Jαk und auf jedem dieser Intervalle Jαi zwei rationale
Zahlen t αi , < tα0i < tαi ; eine Umgebung von x0 = {t10 , t20 , . . . , tα0 , . . .} besteht dann
definitionsgemass aus allen Punkten x = {t1 , t2 , . . . , tα , . . .} die den Bedingungen
tα i < tαi < t"αi genügen. (Tychonoff, [1])
Tychonoff says: Jα is a collection of copies of [0,1], Rτ is the set of points x
with coordinates {t1 , . . . , tα , . . .}, tα ∈ Jα . If we choose finitely many Jα1 , . . . , Jαk
and rational numbers tα i < tα0i < tαi , a neighbourhood of x0 consists, by definition,
of points x of Rτ with tα i < tαi < tαi . Čech [2] points out that the definition carries
over to the general case. The general definition is given thus:
Let {Si } be a family of sets; the subscript runs over an arbitrary given set I . The
cartesian product Pi Si of the family {Si } is the set of all families x = {xi }, each
xi belonging to X i . The xi are called coordinates of x. If every Si is a topological
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 79
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_6
80 6 Topologies Defined by Maps
We now take up ‘initial topology’, the first of the two ways, mentioned above, of
defining topologies using maps.
Proposition 6.1 Let X be a set and let {Yi } be a family of topological spaces. For
each i, let a map f i : X → Yi be given. There is the smallest topology on X making
each f i continuous.
Proof The required topology is the intersection of all topologies making each f i
continuous. This family is nonempty as the discrete topology is in it. Explicitly, the
topology is generated by the family of all sets of the form f i−1 (Vi ) as Vi varies over
all open sets in Yi and i runs through all the indices under consideration.
Basic open sets for initial topologies and continuity for maps from a space with initial
topology are easily described.
Proposition 6.3 Let X be endowed with the initial topology generated by a family
f i : X → Yi of maps. Then
1) the sets of the form ∩n1 f i−1
k
(Uik ), where i 1 , . . . , i n are any finite number of
indices and Uik are open sets in Yik , form a basis for X ;
2) for any topological space Y , a map g : Y → X is continuous if and only if
f i ◦ g : Y → Yi is continuous for all i.
6.1 Initial and Final Topologies 81
Proof 1) is immediate: sets of the form f i−1 (Ui ), Ui ⊂ Yi open, is a subbasis for the
initial topology, by definition, and hence finite intersections of these form a basis.
2) The f i are continuous and so if g is continuous, so are f i ◦ g for all i. Conversely,
suppose all the f i ◦ g are continuous. Then, for any open Ui ⊂ Yi , g −1 ( f i−1 (Ui )) is
open in Y . This means that g pulls back all the subbasic open sets f i−1 (Ui ) to open
sets and so is continuous.
We will look at the important instance of the product topology in the next section.
Here is the dual construction, the ‘reverse method’.
Proposition 6.4 Let Y be a set and let {X i } be a family of topological spaces. For
each i, let a map gi : X i → Y be given. Then there is the largest topology on Y
making each gi continuous.
Proof The indiscrete topology on Y makes all the gi continuous. So the collection
T
of all topologies on Y making all the gi continuous
is nonempty. Define T{gi } :=
{T : T ∈ T }, the topology generated by {T : T ∈ T }. This is the required
topology.
The topology given by the proposition is called the final topology on Y given by
the maps {gi }. An important instance of this topology, the quotient topology, will be
considered later in the chapter.
Proposition 6.5 Let {X i } be a collection of spaces and let a set Y be given the final
topology given by a set of maps gi : X i → Y . Then
1) V is open in Y if and only if gi−1 (V ) is open in X i for every i;
2) for any topological space Z , a map h : Y → Z is continuous if and only if
h ◦ gi : X i → Z is continuous for all i.
Example 6.6 1. Let {Ti } be a collection of topologies on a set X and let idi :
(X, Ti ) → X be the identity
map for all i. Then the final topology on X generated
by the family {idi } is Ti = inf Ti .
2. The function p(x) = e2πı x , x ∈ [0, 1], maps [0,1] onto the unit circle S1 , is one-
one on (0,1) and p(0) = p(1) = 1. The set of fibres p −1 (z) for z ∈ S1 are singletons
{x}, x ∈ (0, 1) for z = 1 and p −1 (1) = {0, 1}. Let X̃ denote the set of fibres: X̃ :=
{ p −1 (z) : z ∈ S1 }. Note that X̃ is got by ‘identifying’ the end points of [0,1]. Give X̃
the final topology defined by the map g : [0, 1] → X̃ , g(x) = x̃ := p −1 ( p(x)). Then
X̃ , a continuous image of the compact space [0,1], is compact. The map h : X̃ →
S1 defined by h(x̃) = p(x) is well-defined and is continuous because h ◦ g = p :
[0, 1] → S1 is continuous. Thus h is a bijective continuous map from the compact
82 6 Topologies Defined by Maps
The last example is a special instance of ‘quotient spaces’ that will be discussed in
the final section of the chapter.
While discussing basic open sets (Chap. 3), we mentioned the product of two topo-
logical spaces. Let us recall that definition. If X, Y are two topological spaces, the
collection B of subsets of X × Y of the form U × V , where U ⊂ X, V ⊂ Y are
open, satisfies the needed conditions to form a basis for a topology on X × Y .
This topology with B as a basis is called the product topology on X × Y . Let
us look at this topology from a different perspective now. Consider the projec-
tion maps p : X × Y → X, q : X × Y → Y given by p(x, y) = x, q(x, y) = y. For
open sets U ⊂ X, V ⊂ Y , p−1 (U ) = U × Y and q−1 (V ) = X × V and so p, q
both pull back open sets to open sets, showing that they are continuous. More-
over, if T is any topology on X × Y such that both p and q are continuous
and U ⊂ X, V ⊂ Y are open, then p−1 (U ) = U × Y, q−1 (V ) = X × V ∈ T , hence
U × V = (U × Y ) ∩ (X × V ) is in T . This means B ⊂ T and so T contains the
product topology. We have proved that the product topology is the smallest topology
on X × Y that makes both the projections p, q continuous. Thus, it is the initial
topology defined on X × Y by the projections p and q. We take cue from this to
define the product topology for any collection of spaces.
To undertake a full fledged discussion of product spaces, which is a desideratum
for a basic understanding of topology, it is essential to first understand the meaning
of the cartesian product set of an arbitrary collection of sets. Let us first look at a
finite product X 1 × X 2 × . . . X n of sets. This, by definition, is the set of n-tuples
(x1 , x2 , . . . , xn ) with x j ∈ X j for each j ∈ {1, 2, . . . , n}. But what is an n-tuple?
Writing x( j) for x j , we see that a we have a map x : {1, 2, . . . , n} → ∪ j X j with
x( j) ∈ X j for each j and the n-tuple is just the listing out of all the values of this
map on the set {1, 2, . . . , n} as (x(1), x(2), . . . , x(n)). With this in mind, we can now
give the definition of the cartesian product of an arbitrary indexed family {X i }i∈I
of nonempty sets. (The index set I will not be usually mentioned explicitly, except
when necessary.) The product X = i X i is, by definition, the set of all maps x on
the index set I taking values in X i with x(i) ∈ X i for every indexi ∈ I. Thus,
for example, a sequence (xn ) in a set X is an element of the product n X n where
X n = X for each n ∈ N. In consonance with the notation in this case, conventionally
one writes x(i) as xi and x = (xi ) ∈ i X i . When all the X i are equal, say X i = X
for all i, we also write X I for X i . With this notation, a sequence in X is an element
of X N .
6.2 Product Topology 83
Definition 6.7 Take the product X = i X i of an indexed family {X i } of nonempty
sets. For each i, the i-th projection pi : X → X i is the map given by pi (x) = xi
taking x = (xi ) ∈ X to its ‘i-th coordinate’. .
The product topology can be described in many ways. We formulate our definition
in terms of the projections, as explained above.
Definition 6.8 For a family {X i } of spaces, the product topology on X i is the
initial topology defined by the pi : X i → X i ; i.e. it is the smallest topology that
makes all the projections pi continuous.
As noted, the product topology was defined by Tychonoff. According to J.L.Kelley,
the formulation given here is due to Bourbaki.
Proposition 6.9 The product topology on the cartesian product X = i X i of an
i , i ∈ I} of topological spaces has as a base the collection B of
indexed family {X
sets of the form pi−1 (Ui ) with Ui ⊂ X i open and the intersection is over finite sets
F ⊂ I.
Proof The product topology has as subbase all sets of the form pi−1 (Ui ), Ui ⊂ X i
open, and finite intersections of these give B.
Remark 6.10 1. In X i , a subset of the form Ui is open if and only if each Ui
is open and Ui = X i for all but finitely many i.
2. Thus, an infinite product of discrete spaces is never discrete.
As a space with initial topology, for a product space continuity of maps into it is
characterised in terms of projections.
Proposition 6.11 A map f : Y → X i , from a space to a product space, is con-
tinuous if and only if its composition with each projection, pi ◦ f : Y → X i , is con-
tinuous.
Proof Recall that the product topology is the initial topology defined by the family
pi of projections and the earlier characterisation of continuity of maps into a space
with initial topology 6.3.
A projection not only pulls back open sets to open sets, it also takes open sets to
open sets. A map which takes any open set (closed set) to an open set (a closed set)
is called an open map (a closed map).
Proposition 6.12 Each projection pi : X i → X i is an open map; that is, it maps
open sets to open sets.
Proof Let U be open in X , x = (xi ) ∈ U and B = j∈F p−1 j (U j ) be a basic open
set with x ∈ B ⊂ U , where F is finite. If i ∈ F, then xi ∈ pi (B) = Ui ⊂ pi (U ); if
i∈/ F, then pi (B) = X i = pi (U ). In any case, pi (U ) is a neighbourhood of xi and
the proof is complete.
84 6 Topologies Defined by Maps
Example 6.14 1. The usual topology on Rn is the same as the product topology and
continuity of f : X → Rn , f (x) = ( f 1 (x), ..., f n (x)), is equivalent to the continuity
of each coordinate function f j .
2. If (X, d), (X , d ) are metric spaces, metrics ρ1 , ρ2 and ρ∞ (analogous to the
metrics d1 , d2 and d∞ on R2 ) were introduced in Chap. 2, Exercises. The topology
given by each of this metric is nothing but the product topology on X × X .
3. If d is a metric on X , then d : X × X → R is continuous.
4. Let {(X
n , dn )} be a countable collection of metric spaces and recall the metric
d on X = X n (see 2.10 (6)). Then the associated metric topology on X is just
the product topology.
We start with a set theoretic observation and use it to realise a quotient space as an
‘identification space’.
6.3 Quotient Topology 85
Corollary 6.20 Any continuous map q of a compact space X onto a metric space
Y is a quotient map. The associated quotient space X̃ is homeomorphic to Y .
The result and the proof are valid if compact sets in Y are closed; e.g. Y is a Hausdorff
space; see 8.10.
Such a map r is called a retraction and the set E is called a retract of X . These
will be studied in some detail in Chap. 14.
Example 6.23 1. In Section 1, as an example of final topology, we saw that the circle
S1 = T := {z ∈ C : |z| = 1}, which is also the one dimensional torus, is a quotient
space of the closed interval [0,1] got by identifying the end points. This also follows
from 6.20 taking q(t) = ex p(2πıt).
2. A compact cylinder is got as a quotient space of a square patch by identifying
corresponding points of one pair of opposite edges. For example, in [0, 1] × [0, 1],
identify (x, 0) with (x, 1) for each x ∈ [0, 1]. The reader is urged to write down the
details, using Fig. 6.1.
6.3 Quotient Topology 87
a2 c2 b2
U2 W2 V2
U1 W1 V1 Ũ W̃ Ṽ
ã b̃
c̃
a1 c1 b1
Fig. 6.1 Cylinder as a quotient of the square: ã = q(a1 ) = q(a2 ); b̃ = q(b1 ) = q(b2 ); c̃ = q(c1 ) =
q(c2 ).
a2 b2 a3
U2 V2 U3
Ṽ b̃
U1 U4
V1
Ũ
a1 b1 a4 ã
Fig. 6.2 Torus as a quotient of the square: ã = q(a j ), b̃ = q(bi ); Ũ is made up of the U j and Ṽ
of the Vi .
X × {1}
V
[0,1]
X × {0}
Exercise 6
c) Hence, if σ is a permutation of I , then X i is homeomorphic to X σ (i) ; e.g.
if I is countably infinite X I X N for any space X .
4. Box topology. If {X i } is a family of
spaces, the product Ui of open subsets
Ui in X i form a base for a topology on X i , called the box topology. This equals
the product topology only for finite products.
a) Take RN with the box topology. Then f : R → RN , f (x) = (x, x, . . . , ) is not
continuous. What about the map g(x) = (x/n)?
b) What is the closure in RN of the set of all sequences whose terms are eventually
0 in i) the product topology, ii) the box topology.?
5. Quotients. a) The projection p1 : R2 → R is a continuous, open surjection,
but is not a closed map.
b) Let E = {(x, y) ∈ R2 : x ≥ 0} ∪ {(x, 0) ∈ R2 : x ∈ R}. Then p1 : E → R is
a quotient map which is neither open nor closed.
c) If f : X → Y is continuous, then f = g ◦ q where q : X → Z is a quotient
map and g : Z → Y is a continuous injection, for a space Z .
d) A subspace of a quotient may not be the quotient of a subspace. Make the
statement precise and justify.
6. Möbius band. Here is a ‘twisted’ version of a (compact) cylinder. Take [0, 1] ×
[0, 1] and identify (x, 0) with (1 − x, 0) for each 0 ≤ x ≤ 1. The resulting quotient is
a ‘one-sided’ surface called the Möbius band, named after the German mathematician
August Ferdinand Möbius. Elaborate on the following picture, Fig. 6.4.
7. Klein bottle. Form the quotient of [0, 1] × [0, 1] by identifying (x, 0) with
(x, 1) and (0, y) with (0, 1 − y) for all 0 ≤ x, y ≤ 1. The resulting space is a ‘twisted
torus’ called the Klein bottle, after the well-known German mathematician Felix
Klein. Draw the square and mark the identifications and try to draw a picture of the
surface.
8. Projective space. On X = Rn+1 \ {0}, define x Ry if 0 lies on the straight line
joining x and y. This is an equivalence relation and the quotient space X/R is called
the projective space RPn .
a) RPn is homeomorphic to the quotient of the sphere Sn got by identifying any
x with its antipodal point −x.
b) RP1 , the ‘projective line’, is homeomorphic to S1 .
9. Suspension. For a space X , the suspension S(X ) is the quotient X × [0, 1]/R
where R is the equivalence relation given by (x, 0)R(y, 0) and (x, 1)R(y, 1) for
x, y ∈ X , (i.e. it is obtained by pinching each of X × 0 and X × 1 to a point (Fig. 6.5).
(Picturise a double cone with base X .) What is the relation between Cone(X ) and
S(X )?
S({0, 1}) is (homeomorphic to) the circle S1 .
10. The upper half-plane. G = S L(2, R), the group of real 2 × 2 matrices of
determinant one, acts on H := {z ∈ C : I m(z) > 0}, the upper half-plane, as Möbius
transformations z → az+b
cz+d
. Check:
6.3 Quotient Topology 91
b2 c2 a2
U2 W2 V2
U1 W1 V1
a1 c1 b1
X × {1}
V1
[0,1]
V0
X × {0}
a) If q is the quotient map, then the restrictions of q to E c and Y are both home-
omorphisms, q(E c ) is open and q(Y ) is closed in X + f Y .
b) If f is the constant map f (E) = y0 , then X + f {y0 } is the space X E obtained
from X by pinching the set E to a point.
c) Cone (X ) is got by attaching X × [0, 1] to a point y0 by the map f (X × {1}) =
y0 .
d) Let X, Y be disjoint compact intervals in R, say X = [0, 1] and Y = [3, 4].
Take E = {0, 1} and define f (0) = 3, f (1) = 4. Then X + f Y is homeomorphic to
the circle S1 .
13. Topological groups—products and quotients. a) The product of any
nonempty collection of topological groups is a topological group.
b) Let G be a topological group and H a subgroup. Equip the left coset space
G/H with the quotient topology. H is closed if and only if singletons are closed in
G/H and H is open if and only if G/H is discrete.
6.4.1 R. L. Moore
Robert Lee Moore (1882–1974) was an American topologist who spent most of his
career at the University of Texas, a student of the famed topologist Oswald Veblen.
He is best known for his unusual teaching method, now called the Moore method.
He had 50 doctoral students, including many well-known topologists.
6.4.2 Möbius
6.4.3 Klein
Felix Klein (1849–1925) was a German mathematician. His doctoral adviser at Bonn
was the geometer Julius Plücker. He was appointed as a Professor at Erlangen when
he was only 23. He moved to München, to Leipzig and, finally, to Göttingen. His
famous Erlangen Programme was initiated in his inaugural lecture at Erlangen and
was completed and published at Göttingen. It presents geometry as the study of the
properties of a space that is invariant under a group of transformations, thus unify-
ing geometry and encompassing both Euclidean and nonEuclidean geometries. The
work had had a profound influence on the development of mathematics and its ideas
have since been ingrained in general mathematical thinking. He sought to re-establish
Göttingen’s primacy in world mathematics and succeeded in bringing David Hilbert
to Göttingen. His notable contributions were in geometry, group theory, invariant the-
ory, number theory, elliptic modular functions, automorphic functions and reforms to
mathematical education in secondary schools. Klein bottle, his nonorientable surface,
cannot be embedded in the three-dimensional Euclidean space.
References
1. Tychonoff, A.: Über die Topologische Erweiterung von Räumen. Math. Ann. 102, 544–561
(1930). https://doi.org/10.1007/BF01782364
2. Čech, E.: On bicompact spaces. Ann. Math 2(38), 823–844 (1937). https://doi.org/10.2307/
1968839
3. Moore, R.L.: Concerning upper semi-continuous collections of continua. Trans. Am. Math. Soc.
27, 416–428 (1925). https://doi.org/10.2307/1989234
Chapter 7
Products of Compact Spaces
Tychonoff’s product theorem is one of the most useful and important results of
topology. Here is how Tychonoff stated it in his celebrated paper.
Betrachten man unter den Elementen ... nur diejenigen Funktlonen f(x), die der
Ungleichung −k ≤ f (x) ≤ k, 0 ≤ x ≤ 1, für eine feste natürliche Zahl k genügen,
so erhält man eine Teilmenge Fk ... , die ihrerseits als topologischer Raum (als Rela-
tivraum ...) betrachtet werden kann. Von diesen Räumen Fk beweisen wir nun, daßsie
bikompakte Hausdorffscke Räume sind. (Tychonoff, 1930, [1]).
He says that ‘the subset of those functions f such that −k ≤ f (x) ≤ k, 0 ≤ x ≤ 1,
for a fixed natural number k, can be viewed as a topological space (relative subspace).
We now show that these spaces are bicompact, Hausdorff.’ In modern notation, this
says that the product space [−k, k][0,1] is compact (and Hausdorff). Later, Tychonoff
remarked that his proof was valid for products of arbitrary compact spaces.
Recall (Chap. 6) that Tychonoff defined the product topology only on [0, 1][0,1]
in [2] and this was carried over to products of arbitrary spaces by Čech, [3]. Čech
also gave the proof of the general form of Tychonoff’s theorem. He stated the result
simply as follows.
The Cartesian product S = Pi Si of any family of bicompact spaces is a bicompact
space.
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 95
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_7
96 7 Products of Compact Spaces
Proof (via open covers) First we set up some needed notation. Let X = i∈I X i ;
recall that it is the set of all maps x with domain I such that x(i) ∈ X i for i ∈ I (See
appendix below). For J ⊂ I, we call X J = j∈J X j a partial product. An element
u = u J ∈ X J is a map with domain J such that u( j) ∈ X j for j ∈ J and will be
called a partial map. If u, u
are partial maps with disjoint domains J, J
, then write
u u
∈ X J × X J
for the partial map with domain J ∪ J
whose restrictions to
J, J
are, respectively, u, u
. Then x ∈ X may be identified with x| J x| I \J . Note
that elements of X J are precisely the restrictions x| J , x ∈ X = X I and x → x J is
just the projection p J : X → X J . Observe also that when J = ∅, the only partial
map is the empty map and so X ∅ = {∅}.
We can now begin the proof. Suppose X is not compact and {Ua }a∈A is an open
cover of X with no finite subcover. Let P be the set of all partial maps u such that
for any neighbourhood V J of u in X J , the pull back p−1 J (V J ) cannot be covered by
finitely many Ua , where J = domain u. Note that the assumption on the open cover
means that the empty partial map ∅ ∈ P. To complete the proof, we show that there
is an x ∈ P with domain I , i.e. the partial map x is a full map: x ∈ X . We get this x
by applying Zorn’s lemma to P.
Step 1: Partial order on P. Define the partial order as the natural extension of
maps: u u
if domain u = J ⊂ J
= domain u
and u = u
| J . In this step, we show
that if u ∈ P and v u, then v ∈ P. Write domain u = J and domain v = K . If
V is a neighbourhood of v in X K and if q = q J K is the projection X J → X K , then
q−1 (V ) is a neighbourhood of u in X J . Since u ∈ P, p−1 −1
J (q (V )) cannot be covered
by a finite number of Ua . But p K = q ◦ p J and so p J (q (V )) = p−1
−1 −1
K (V ) and we
conclude v ∈ P.
Step 2: In this step, we apply Zorn’s lemma to get a maximal element in P. To
ensure the applicability of Zorn’s lemma, we prove that any chain C in P has an upper
bound in P. To get an upper bound, take J0 to be the union of the domains of all partial
maps u ∈ C. Then u 0 on J0 given by u 0 ( j) = u( j) if j ∈ J =domain u, u ∈ C, is well
defined as C is a chain. To see that u 0 ∈ P it is enough to check finitely many Ua do
not cover p−1 n −1
J0 (U ) for any basic neighbourhood U of u 0 in X J0 , U = ∩1 p jk (W jk ),
jk ∈ J0 , 1 ≤ k ≤ n, W jk ⊂ X jk open. Write F = { j1 , · · · , jk } and u F = u| F . Since
F is finite and C is a chain, there is a v0 ∈ C with u F v0 . But then, by Step 1,
u F ∈ P. Note that U = q−1 J0 F (W ) where W = k W jk is a neighbourhood of u F in
X F . Hence p−1 F (W ) = p −1
J0 (U ) cannot be covered by finitely many Ua . Thus u 0 ∈ P,
C has an upper bound in P and Zorn’s lemma yields a maximal element x0 ; let
domain x0 = I0 .
Step 3: We complete the proof by proving that I0 = I and x0 ∈ X .
Assume there is an i ∈ I \ I0 . For a u ∈ X i consider the partial map x0 u. Write
x0u = x0 u and domain x0u = I0i = I0 ∪ {i}. We claim that x0u ∈ P for some u ∈ X i .
Suppose not. Then, for each u ∈ X i , there is a neighbourhood Vu = Wu × Nu of x0u
98 7 Products of Compact Spaces
can be covered by finitely many Ua since each set on the right can be thus covered.
This contradicts the maximality and shows that I0 = I. Thus x0 ∈ X and so x0 ∈ Ua
for some a. Since p I = id X , p−1
I (Ua ) = Ua is covered by a single Ua , a contradiction
as x0 ∈ P. Thus the assumption that {Ua }a∈A has no finite subcover is untenable.
A simple proof of Tychonoff’s theorem using nets was given by Paul Chernoff
[5] in 1992. We now give a variation of this proof, following Kenneth Ross [6]. The
concluding, clinching part of this proof is separated as a little lemma due to Charles
Pugh (2003), see [6].
Lemma 7.4 Let X be any topological space and Y a compact space. Let {(xα , yα )}α∈D
be a net in X × Y . If (xα ) has a cluster point x in X , then (x, y) is a cluster point of
{(xα , yα )} for some y ∈ Y .
Proof Let D := {(α, Ux ) : α ∈ D, Ux ∈ Ux , xα ∈ Ux }. It is a directed set with
order defined by (α, Ux ) (β, Vx ) if α ≤ β and Vx ⊂ Ux . Define η : D → Y by
η(α, Ux ) = yα . This is a net in the compact space Y , so has a cluster point y ∈ Y .
We prove that (x, y) is a cluster point of the net {(xα , yα )}. For this, take a basic
neighbourhood U × V of (x, y). Since x is a cluster point of (xα ), for any α ∈ D
there is a β ≥ α such that xβ ∈ U and since y is a cluster point of η, there is a
(γ , Vx ) (β, U ) with yγ = η(γ , Vx ) ∈ V .This means that γ ≥ α and (xγ , yγ ) ∈
U × V proving that (x, y) is a cluster point of the net {(xα , yα )}.
Proof (Tychonoff’s Theorem via nets.)
Let X = i∈I X i be a product of compact spaces. It suffices to show that every
net (xα ) in X has a cluster point.
A partial net of (xα ) is a net of the form (xα | J ) in X J for some J ⊂ I . A cluster
point of a partial net is a partial cluster point of (xα ). When J = { j}, the partial net
(xα | J ) is a net in X j and, by compactness, has a cluster point. Hence the set P of all
partial cluster points of (xα ) is nonempty. P has a natural order: x J ≤ x K if J ⊂ K
and x K | J = x J . Our strategy is to get a maximal partial cluster point in P via Zorn’s
lemma and prove that it is really a cluster point.
To be able to invoke Zorn’s lemma, we show that every totally ordered subset
T ⊂ P has an upper bound in P. To do that, let J0 = ∪{J : x J ∈ T } and define
y0 ( j) = x J ( j) for j ∈ J with x J ∈ T . This is well-defined since T is totally ordered.
We have to show that y0 ∈ P, i.e. that y0 is a cluster point of (xα | J0 ). For this, take
a basic neighbourhood U0 of y0 in X J0 , U0 = ∩ j∈F p−1 j (U j ) where F ⊂ J is finite
and U j is open in X j for j ∈ F. As T is totally ordered, there is an x K ∈ T with
F ⊂ K . Since x K is a cluster point of (xα | K ), for any β there is an α ≥ β such that
xα ( j) ∈ U j for j ∈ F. This means xα | J0 ∈ U0 . We have thus proved that the xα | J0
7.1 Tychonoff’s Theorem 99
Example 7.5 1. Products of copies of [0,1] (‘cubes’) are compact. In particular, the
infinite dimensional cube [0, 1]N is compact.
2. Infinite products of finite discrete spaces are compact (and nondiscrete!). For
example {0, 1}N is compact. Observe that {0, 1}N is the space of sequences whose
terms are either 0 or 1.
The Axiom of Choice was used in both the proofs, in the form of Zorn’s lemma.
It is unavoidable. In fact, S.Kakutani conjectured in 1935 and Kelley proved in 1950
[7] that Tychonoff’s theorem is equivalent to the Axiom of Choice. Here is Kelley’s
simple proof.
Proof Assume Tychonoff’s theorem. We prove the Axiom of Choice in one of its
equivalent formulations: the product of any nonempty collection of nonempty sets
is nonempty.
So let {Si : i ∈ I } be such a collection. When the index set is finite, the choice
axiom is not needed, so we assume this case and suppose that I is infinite. Let X i be
the disjoint union X i = Si {i}. Equip X i with the topology Ti = Tco f ∪ {i}, i.e. the
set {i}.
open sets in X i are the empty set, complements of finite sets and the one-point
Each X i is compact and Si is closed in X i , hence so are pi−1 (Si ) in X = X i , where
pi are the projections. We show: {pi−1 (Si )} has finite intersection property. Let F be
a finite subset of I . Take s j ∈ S j for j ∈ F (No Axiom of Choice!). Define x ∈ X
by x( j) = s j for j ∈ F and x(i) = i for i ∈ I \ F. Then x ∈ ∩ j∈F p−1 j (S j ). Thus
the family {pi−1 (Si )} of closed sets in the compact space X has finite intersection
property and hence ∩i∈I pi−1 (Si ) is nonempty. But i∈I Si = ∩i∈I pi−1 (Si ) and the
proof is complete.
of C are arbitrary nonempty sets, can we still get such a function? Does such a
function exist? Such a function, if it exists, is called a choice function. If each C ∈ C
is well-ordered, then we can define χ as before. Otherwise, it is not clear that a
choice function always exists. One may feel that since each C is nonempty, one can
‘choose’ an element n C ∈ C and define χ as before. But is there a way of ‘choosing’
an n C for each C ∈ C ? As observed earlier, if every set can be well-ordered, then our
argument works. However, the statement that every nonempty set can be well-ordered
is equivalent to the existence of a choice function.
The edifice of modern mathematics rests on the foundations of set theory initiated
by the work of Cantor and mathematical logic. It was realised that indiscriminate use
of set theory led to contradictions and paradoxes. These logical inconsistencies led
to an axiomatic development of set theory. There are several axiomatic approaches:
Zermelo–Fraenkel theory [8, 9]; see also [10], Hilbert–Bernays–von Neumann sys-
tem, etc. Most mathematicians take a naive point of view of set theory and we do
the same in this book. We are not going to present any axiomatic treatment of set
theory. (See Halmos, [11] for a nice, accessible treatment.) In fact, we are not going
to discuss set theory at all. Basic concepts, standard notation and results of ‘set
theory for a working mathematician’ are all assumed. In the proof of Tychonoff’s
theorem, Axiom of Choice is used in its equivalent form of Zorn’s lemma, and con-
versely, Tychonoff’s theorem implies the Axiom of Choice. Our purpose here is to
merely mention some of the related formulations of one of the most famous axioms
in modern mathematics.
A partial order on a nonempty set X is a relation that is reflexive, transitive and
antisymmetric. A set with a partial order is called a partially ordered set. A partial
order is a total order if for any two x, y ∈ X , either x y or y x. A totally ordered
set is also called a chain. A partially ordered set X is well-ordered if every nonempty
subset of X has a least element. An element M of a partially ordered set X is maximal
if M x ∈ X implies M = x. A minimal element is defined analogously. An upper
bound for A ⊂ X is an element u ∈ X such that a u for all a ∈ A.
Examples: R, with the usual order, is totally ordered, but not well-ordered. The
same is true for Q and Z, but N is well-ordered. The set P(X ) of all subsets of a set
X is partially ordered, but not totally ordered, by set inclusion.
Zorn’s Lemma: If X is a partially ordered set such that every totally ordered
subset has an upper bound in X, then X has a maximal element.
Axiom of Choice: For every nonempty collection of nonempty sets, there is a
choice function.
Theorem: The following are equivalent:
1) Axiom of Choice
2) Given a collection C of disjoint nonempty sets, there is a set A consisting of
exactly one element from each C ∈ C . (Zermelo Postulate)
3) The Cartesian product of any nonempty collection of nonempty sets is nonempty.
4) If E is a collection of sets and C is a chain in E , then there is a maximal chain
M in E containing C . (Hausdorff Maximal Principle)
5) On any nonempty set X, there is a well-ordering. (Well-ordering Principle)
7.2 Appendix: Axiom of Choice 101
See Kelley [12] or Moore [13], for example. Kelley takes the Hausdorff maximal
principle as an axiom and proves the others. Of these, Zorn’s lemma is the most
commonly used formulation these days. We have used it in the proof of Tychonoff’s
theorem. For its use in algebra, see the exercises. Kuratowski (1922) and Salomon
Bochner (1922) had given versions of Zorn’s lemma much before Max Zorn (1935).
Axiom of Choice and its equivalents have a tangled history. It has a large number
of equivalents in various fields. There are whole books devoted exclusively to such
matters, e.g. G.H Moore’s tome, [13].
One final word. The place of Axiom of Choice in set theory had been the object
of long, vigorous study. Kurt Gödel showed (1938) that the negation of the Axiom is
consistent with Zermelo–Fräenkel set theory. Paul Cohen (Fields medallist) proved
(1963) that Axiom of Choice itself is consistent with Zermelo–Fräenkel set theory.
Thus Axiom of Choice is independent of Zermelo–Fräenkel set theory, [14, 15].
Exercise 7
d) Apply the preceding considerations to a product X = X i of compact spaces,
the subbase S consisting of sets of the form pi−1 (Vi ) (pi being the projections and
Vi open in X i ) and the associated base to get a proof of Tychonoff’s theorem.
5. Tychonoff using f.i.p.(Bourbaki) a) Let X be a nonempty set and let F be a
collection of subsets of X with f.i.p. Then there is a family E ⊃ F with f.i.p. property
and satisfying the following conditions: i) if E ∈ E and F ⊃ E, then F ∈ E; ii) if
E 1 , E 2 ∈ E then E 1 ∩ E 2 ∈ E; iii) if a subset E 0 of X intersects every set in E, then
E 0 ∈ E.(P be the family of all collections of subsets of X with f.i.p. that contain F,
partially ordered by inclusion. Apply Zorn’s lemma to get a maximal collection E;
this collection
satisfies the requirements.)
b) { Ē : |E ∈ E} is nonempty, hence so is { F̄ : F ∈ F}, completing the proof
of Tychonoffs theorem.
6. Alaoglu Theorem. a) Let X be a normed space with closed unit ball B.
Let B ∗ denote the set of scalar-valued linear maps f on X satisfying the property
| f (x)| ≤ x for all x ∈ X . If D is the closed unit disc in the complex plane, then
B B∗ = { f | B : f ∈ B ∗ } is a subset of D B since | f (x)| ≤ 1 for x ∈ B and f ∈ B ∗ . It
is a closed subset of the product D B and hence is compact by Tychonoff’s theorem.
b) Suppose X as in a) is also an algebra with identity satisfying x y ≤ xy
for all x, y ∈ X (so X is normed algebra). Then the set (X ) := { f | B : f ∈
B ∗ , f (x y) = f (x) f (y)∀x, y ∈ X } ⊂ D B is also compact. (Actually, for any linear
functional f on X satisfying the condition f (x y) = f (x) f (y) for all x, y (multi-
plicative linear functional), the condition f ∈ B ∗ is automatic. The topology of B ∗
as a subspace of D B is called the weak ∗ topology in functional analysis.)
7.3.1 Bourbaki
out definitive tomes on all of pure mathematics. The group met periodically and had
fierce, heated, thorough discussions on each one’s draft on the subject under consid-
eration until unanimous agreement was reached for the final draft. Armand Borel,
a later Bourbaki talks of ‘a truly unselfish, anonymous, demanding work by people
striving to give the best possible exposition of basic mathematics’ and says ‘the activ-
ity within Bourbaki was a tremendous education, a unique training ground, obviously
a main source of the breadth and sharpness of understanding... The requirement to
be interested in all topics clearly led to a broadening of horizon’, [17]. (Quoted with
the permission of the American Mathematical Society.) These sentiments are shared
by many members of the group (e.g. Dieudonné [18] and Cartan [19], p xix). The
proposed series was called Éléments de Mathématique and volumes on Theory of
Sets, General Topology, Real Analysis, Algebra, Commutative Algebra, Topological
Vector Spaces, Spectral Theory, Integration, Lie Groups and Lie Algebras, History of
Mathematics and, after a gap of several decades, Algebraic Topology have appeared
so far. The group is ‘dynamic’: older members are retired and new, younger mem-
bers are drafted. The books are known for their rigour, generality and emphasis on
mathematical structure. They have had considerable influence on the development
of mathematics. In its periodic seminars, Séminaires Bourbaki, significant recent
developments are discussed by experts and the proceedings are published.
7.3.2 Čech
Eduard Čech (1893–1960) was a Czech mathematician known for his work on
topology and geometry. Čech cohomology and Stone-Čech compactification are
well-known. The first published proof (1937) of the general form of Tychonoff’s
theorem is due to him.
7.3.3 Tychonoff
for differential equations and proved uniqueness results for the heat equation. His
work on ill posed problems is also well-known.
7.3.4 Kelley
John Kelley (1916–1999) was an American mathematician who spent most of his
academic life at Berkeley. He mainly worked in topology and functional analysis.
The concept of a subnet was introduced by him in 1950 and, in fact, the term ‘net’
itself was coined by him. His 1955 book General Topology, one of the earliest in
English, was a classic. Its Appendix on set theory is an oft-cited short and succinct
presentation of the foundations, now called Morse–Kelley Set Theory. He also proved
the equivalence of the Tychonoff product theorem and the Axiom of Choice. His
other books include Linear Topological Spaces (with I. Namioka) and Measure and
Integral (with T.P.Srinivasan). He was visiting the Indian Institute of Technology,
Kanpur, India, (the current author’s alma mater) during the formative years of the
Institute. This author first learned the subject from Kelley’s book. (The course was
taught by the author’s teacher at Kanpur, U.B.Tewari, an alumnus of UC, Berkeley.)
Solving the exercises was as much excitement as reading the text was a pleasure.
References
1. Tychonoff, A.: Über einen Funktionenraum. Math. Ann. 111, 762–766 (1935). https://doi.org/
10.1007/BF01472255
2. Tychonoff, A.: Über die Topologische Erweiterung von Räumen. Math. Ann. 102, 544–561
(1930). https://doi.org/10.1007/BF01782364
3. Čech, E.: On bicompact spaces. Ann. Math 2(38), 823–844 (1937). https://doi.org/10.2307/
1968839
4. Matheron, E.: Three proofs of Tychonoff’s theorem. Am. Math. Monthly 127, 437–443 (2020).
https://doi.org/10.1080/00029890.2020.1718951
5. Chernoff, P.R.: A simple proof of Tychonoff’s theorem via nets. Am. Math. Monthly 99, 932–
934 (1992). https://doi.org/10.1080/00029890.1992.11995956
6. Ross, K.A.: Appendix to Informal Introduction to Set Theory (2003). https://pages.uoregon.
edu/math/people/ross/SetTheoryAppendix.pdf
7. Kelley, J.L.: The Tychonoff product theorem implies the axiom of choice. Fund. Math. 37,
75–76 (1950). https://doi.org/10.4064/fm-37-1-75-76
8. Zermelo, E.: Untersuchungen über die Grundlagen der Mengenlehre I. Math. Ann. 65, 261–281
(1908). https://doi.org/10.1007/BF01449999
9. Fränkel, A.: Zu den Grundlagen der Cantor-Zermeloschen Mengenlehre. Math. Ann. 86, 230–
237 (1922). https://doi.org/10.1007/BF01457986
10. Fränkel, A.: Abstract Set Theory. North-Holland (1953)
11. Halmos, P.R.: Naïve Set Theory. Affiliated East-West Press (1960)
12. Kelley, J.L.: General Topology. Affiliated East-West Press (1969)
13. Moore, E.H.: General Analysis I. Part II. Mem American Philosphical, Society (1939)
14. Cohen, P.J.: Independence Results in Set Theory, Theory of Models, Proceedings of Sympo-
sium, Berkeley, 1963, North Holland, pp. 39–54 (1965)
References 105
15. Gödel, K.: The consistency of the axiom of choice and the generalized continuum hypothesis
with the axioms of set theory. Uspehi Mat Nauk (N.S.) 3, 96–149 (1948)
16. Weil, A.: The Apprenticeship of a Mathematician (Translated from the French by J. Gage.)
Birkhäser (1992)
17. Borel, A.: Twenty-Five Years with Nicolas Bourbaki (1949–1973). Not. Am. Math. Soc. 45,
373–380 (1998)
18. Dieudonné, J.A.: Une Généralisations des Espaces Compacts. J. Math. Pures. Appl. 23, 65–76
(1944)
19. Cartan, H.: Oeuvres, vol. 1. Springer (1979)
Chapter 8
Separation Axioms
‘Separation’ here refers to separation of points and closed sets by open sets. Haus-
dorff, [1], states the separation axioms as follows:
Trennungsaxiome.
Ist x1 = x2 , so gibt es eine offene Menge G 1 die x1 , aber nicht x2 enthält.
Ist x1 = x2 , so gibt es zwei disjunkte offene Mengen G 1 , G 2 mit x1 ∈ G 1 , x2 ∈ G 2 .
Ist F2 eine den Punkt x1 nicht enthaltende abgeschlossene Menge, so gibt es zwei
disjunkte offene Mengen G 1 , G 2 mit x1 ∈ G 1 , F2 ⊂ G 1 .
Sind F1 , F2 zwei disjunkte abgeschlossene Mengen, so gibt es zwei disjunkte
offene Mengen G 1 , G 2 mit F1 ⊂ G 1 , F2 ⊂ G 2 .
(‘Separation axioms:
If x1 = x2 , there is an open neighbourhood G 1 of x1 not containing x2 .
If x1 = x2 , there are disjoint open sets G 1 , G 2 with x1 ∈ G 1 , x2 ∈ G 2 .
If F2 is a closed set not containing x1 , there are two disjoint open sets G 1 , G 2
with x1 ∈ G 1 , F2 ⊂ G 2 .
If F1 , F2 are disjoint closed sets, there are two disjoint open sets G 1 , G 2 with
F1 ⊂ G 1 , F2 ⊂ G 2 .’)
Alexandroff-Hopf [2] call these: Fréchetsches, Hausdorffsches, Vietorissches and
Tietzesches Trennungsaxiome; these are now called T1 , T2 (Hausdorff), T3 (regularity)
and T4 (normality) axioms, respectively. Regularity and normality were defined by
Vietoris [3] and Tietze [4], respectively. T1 -spaces (Riesz spaces or Fréchet spaces)
were first considered by Riesz [5] and by Fréchet [6]. The German term Trennungsax-
iome, meaning ‘separation axioms’, was first used by Tietze [4] in 1923.
Separation axioms Tk , for each k ∈ {0, 1, 2, 3, 4, 5, 3 12 } have been studied (you
know where the T comes from!). We discuss only some of the important ones. A
word of warning on terminology: the use of regular, normal etc. vary considerably
in different references. For instance, in Kelley [7] normality means separation of
closed sets and T4 means normal+T1 . The reverse usage can also be seen in some
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 107
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_8
108 8 Separation Axioms
references. Similar remarks apply to regularity and complete regularity. So, while
looking at any reference the reader should check the usage.
We begin with the separation axiom that was given by Hausdorff initially as a part of
the definition of a topological space. For most purposes in analysis and differential
geometry, Hausdorff spaces are the most basic.
Definition 8.1 A space X is called a Hausdorff space or a T2 -space if distinct points
can be separated by disjoint open sets, i.e. if x, y ∈ X, x = y, there are disjoint open
sets U, V with x ∈ U , y ∈ V .
Proof Let (X, d) be a metric space and x, y ∈ X, x = y. Then the open balls B(x; ε)
and B(y; ε) are disjoint if ε < d(x, y)/2.
Proof 1) ⇒ 2): If x = y, then there are disjoint open sets U, V such that x ∈ U, y ∈
V . Then U ⊂ V c , so Ū ⊂ V c and hence y ∈ / Ū .
2) ⇒ 3): If y = x, 2) says that there is a neighbourhood U of x such that y ∈ / Ū
and so y is not in the intersection mentioned in 3).
3) ⇒ 1): Let x, y ∈ X, x = y. By 3) there is an open set U with x ∈ U but y ∈ / Ū .
So V = Ū c is open, is disjoint from U and y ∈ V .
1) ⇒ 4): If x = y, disjoint open sets U, V separate them, so no net can converge
to both the points, as it cannot be eventually in both.
4) ⇒ 5): If is not closed, there is a net {(xi , xi )} converging to (x, y) with
x = y. But then {xi } converges to both x and y.
5) ⇒ 1): If x = y, then (x, y) ∈
/ . So, if is closed, then are open sets U, V in
X such that (x, y) ∈ U × V ⊂ c . But this means that U, V are disjoint neighbour-
hoods of x, y respectively.
8.1 Hausdorff Spaces 109
The property that one-point sets are closed is equivalent to the T1 axiom, the first in
the list above.
Example 8.4 1. In an infinite set X with the cofinite topology, finite sets are closed,
but X is not Hausdorff; in fact, no two nonempty open sets in X are disjoint. But it
is T1 .
2. Z with arithmetic progression topology is Hausdorff.
Proposition 8.5 In a Hausdorff space, one-point sets are closed and so the space is
T1 .
Proposition 8.6 In a Hausdorff space, a point and a compact set can be separated
by disjoint open sets. That is, if K is a compact subset of a Hausdorff space X and
if x ∈ X , x ∈
/ K , then there are disjoint open sets V, W with x ∈ V, K ⊂ W .
In Hausdorff spaces, compact sets behave like points. We will see many examples
of this. The last result and the next two are instances.
The next result is essentially due to Vietoris, see 8.15. Notice how we are able to
pass from separation of points to separation of a point and a compact set, thence to
separation of compact sets.
Proposition 8.8 In a Hausdorff space, disjoint compact sets can be separated by dis-
joint open sets. In other words, if K 1 , K 2 are disjoint compact subsets of a Hausdorff
space X , there are disjoint open sets V1 , V2 such that K j ⊂ V j , j = 1, 2.
Proof 1) If {xi } is a net converging to x in X and if f (xi ) = g(xi ) for all indices i,
then f (x) = lim f (xi ) = lim g(xi ) = g(x). Note that we have used uniqueness of
limits in a Hausdorff space.
2) is a consequence of 1).
3) If {xi } is a net such that (xi , f (xi )) converges to (x, y) in X × Y , then {xi }
converges to x in X and { f (xi )} converges to y in Y . But continuity of f implies
f (xi ) → f (x) and so, by uniqueness of limits, y = f (x) and the proof is complete.
Proof Since compact sets in a Hausdorff space are closed, repeat the proof of 5.25.
Hausdorff spaces are well behaved vis à vis subspaces and products.
What is the normal meaning of the word ‘normal’? Ordinary, usual, common, stan-
dard are some of the synonyms that come to mind. However, in mathematics, nor-
mally’normal’ means ‘not normal’, not ordinary, something special or abnormal.
Normal operators on Hilbert spaces, normal subgroups and normal topological spaces
are examples. (Having said that, one can also argue that these are the ones that ‘nor-
mally’ arise in practice!)
Let us begin by observing some special properties for disjoint closed sets in
metric spaces. Recall the continuous distance function d(x, E) for any subset E of a
metric space X . d(x, E) = 0 precisely when x ∈ Ē. If A, B are disjoint and closed,
8.2 Normal Spaces 111
The first property is separation of closed sets by continuous functions and the
second one is separation of closed sets by open sets. What we have observed is
that they are equivalent in metric spaces. In general, it is clear that the latter is a
consequence of the former. In fact, it turns out that the two properties are equivalent
in a general setting, but the reverse implication is deep, as we will see. Note also that,
by considering bd(x,A)+ad(x,B)
d(x,A)+d(x,B)
, we can replace 0 and 1 by any two real numbers a, b.
Separation of closed sets is stronger than the separation points in a T1 -space. This
explains the T1 assumption in the next definition. This separation axiom was given
by Tietze in 1923 [8].
The initial discussion, which motivated this definition, shows that metric spaces
are normal.
Proposition 8.14 The following are equivalent for a T1 space X :
1) X is normal.
2) If A is closed and U is open in X with A ⊂ U , then there is an open set V
such that A ⊂ V ⊂ V̄ ⊂ U . (‘Every neighbourhood of a closed set contains a closed
neighbourhood.’)
3) For each pair A, B of closed sets in X , there is a pair of open neighbourhoods
V, W such that V̄ , W̄ are disjoint.
(−a, a)
R
A
R
Here is another large class of normal spaces, besides metric spaces. Vietoris proved
this (even before normal spaces were defined!), [3], Satz 27. It was also later obtained
by Alexandroff and Urysohn [9].
Proof Since closed sets in a compact space are compact, this is a consequence of
the result (8.8) that in a Hausdorff space, disjoint compact sets can be separated.
Example 8.16 Product of two normal spaces may not be normal. R is normal,
but R × R is not. To see the latter assertion, first note that the subspace A =
{(x, −x) : x ∈ R} is an uncountable discrete subspace of R × R (Fig. 8.1). D =
Q × Q is a countable dense subspace. Suppose R × R is normal. For each S ⊂ A,
S and A \ S are disjoint closed sets and so there are disjoint open sets U S , VS with
S ⊂ U S , A \ S ⊂ VS . Since D is dense, D ∩ U S is nonempty. Note that U S ∩ VT is
open and nonempty if S = T . Since D is dense D ∩ U S ∩ VT is nonempty, whereas
D ∩ UT ∩ VT is empty. It follows that D ∩ U S = D ∩ UT if S = T . This means that
S → U S is an injective map from P(A) to P(D). But this is impossible because A
is uncountable whereas D is countable. Thus R × R is not normal.
Remark 8.17 Subspaces of normal spaces may not be normal. [0, 1][0,1] is compact
Hausdorff and hence is normal. But it is a fact that the subspace (0, 1)[0,1] is not
normal. Thus a subspace of a normal space may not be normal. For an example
involving ordinals, see [7, 10, 11].
Proof Let X be normal and C ⊂ X be closed. One-point sets are closed in X and
so are closed in C. If A, B are disjoint closed sets in C, they are also closed in X . If
U, V are disjoint open sets in X separating A, B, then U ∩ C, V ∩ C are open in C
and separate them.
The deepest, fundamental results on normal spaces are Tietze’s result on extension
of continuous functions and Urysohn’s characterisation of normality by separation
of closed sets by continuous functions. We take up the latter result first. We have
seen an easy proof of this for metric spaces in the opening discussion of this section.
As observed earlier, separation of closed sets by continuous functions yield sepa-
ration by open sets. Hence the converse of the following theorem holds for T1 -spaces.
The theorem, obtained by Urysohn [12], is one of the fundamentally important results
of general topology.
Theorem 8.19 (Urysohn’s Lemma, 1925)
Let X be a normal space and A, B be disjoint closed subsets. Then there is a continu-
ous function f : X → [0, 1] such that f (x) = 0 for x ∈ A and f (x) = 1 for x ∈ B.
inf{q ∈ DQ : x ∈ Uq }, x ∈ U1 = B c
f (x) =
1, x∈B
Remark 8.20 There is nothing special about the values 0 and 1. They can be
replaced by any two distinct real numbers, say a < b. In fact, defining f a,b (x) =
(b − a) f (x) + a, we get a continuous function f a,b : X → [a, b] taking values a on
A and b on B. The reader should also check whether T1 axiom was used in the proof.
a b
Proof The case when f is a constant is trivial, so assume that f is not a constant. First
consider the case when f is bounded, with range in [a, b]. Taking c = sup{| f (x)| :
x ∈ E} and composing with a homeomorphism [a, b] → [−c, c], we may replace
[a, b] by [−c, c].
We will make repeated use of Urysohn’s lemma (8.19) and construct the sought
after function as the uniform limit of continuous functions.
Let c0 := sup{| f (x)| : x ∈ E} and consider the disjoint closed sets
A0 = {x ∈ E : f (x) ≤ −c0 /3}, B0 = {x ∈ E : f (x) ≥ c0 /3}.
Urysohn’s lemma yields a continuous function g0 : X → [−c0 /3, c0 /3] with g0 =
−c0 /3 on A0 and g0 = c0 /3 on B0 . Note that |g0 | ≤ c0 /3 and | f − g0 | ≤ 2c0 /3
on E. Repeat this argument with f replaced by f 1 := f − g0 on E, i.e. take c1 =
sup{| f 1 (x)| : x ∈ E}, note that c1 ≤ 2c0 /3 and apply Urysohn’s lemma to the disjoint
closed sets
A1 = {x ∈ E : f (x) ≤ −2c0 /32 }, B1 = {x ∈ E : f (x) ≥ 2c0 /32 }
to get a continuous function g1 : X → [−2c0 /32 , 2c0 /32 ] such that g1 = −2c0 /32
on A1 , g1 = 2c0 /32 on B1 . Take f 2 := f − g0 − g1 and note that | f 1 (x)| ≤ 22 c0 /32
for x ∈ E. Induction gives sequences of continuous functions { f n }∞ ∞
1 on E and {gn }0
on X satisfying
|gn | ≤ 2n c0 /3n+1 and | f n | ≤ 2n+1 c0 /3n+1 , f n := f − n0 gk on E.
Weierstrass M-test shows that ∞ 0 gn converges absolutely and uniformly to a
continuous function g on X . On E, f − g = lim f n = 0 uniformly since | f n | ≤
2n+1 c0 /3n+1 → 0. Moreover
|g| ≤ |gn | ≤ c0 /3 ∞ 0 (2/3) = c0 .
n
The case when f is unbounded reduces to the bounded case on composing with a
homeomorphism R → (−c, c) (e.g. x → 1+|x|
cx
), .
116 8 Separation Axioms
Remark 8.22 The theorem easily extends to complex valued functions and functions
taking values in Rn .
Regularity was defined by Vietoris in 1921 [3] as follows. (The name ‘regular’ was
introduced later by Alexandroff.)
(E) Eine Umgebung Ux eines Punktes x enthalt immer eine Umgebnng Wx von
x, so daßjeder Punkt von CUx (Komplementgrmenge von Ux ) samt einer seiner
Umgebungen in C Wx liegt. (Vietoris [3])
(‘(E) A neighbourhood Ux of a point x always contains a neighbourhood Wx of
x, so that each point of Uxc including one of its neighbourhoods lies in Wxc .’)
The slightly different current definition was introduced by Tietze, [8]. Regular
spaces lie ‘between’ Hausdorff and normal spaces: separation of a point and a closed
set.
Proof 1) Let X be a regular space and let E be a subspace. Let C be a closed set
in E so that C = E ∩ A where A is closed in X . If x ∈ E \ C, then x ∈ / A and by
regularity of X , there are disjoint open sets U, V in X such that x ∈ U, A ⊂ V . Then
8.3 Regular Spaces 117
Separation of disjoint closed sets by disjoint open sets and by continuous functions
turned out to be equivalent. So it is tempting to believe that an analogous result holds
for separation of a point and a closed set, i.e. in a regular space a closed set E and
an x ∈/ E can be separated by continuous functions. But this is not true. Separation
by continuous functions is stronger than separation by open sets in this case. This
necessitates the introduction of a new class of spaces.
Definition 8.27 A T1 space X is said to be completely regular or a Tychonoff space
if for a closed set E in X and a point x0 ∈ X, x0 ∈
/ E, there is a continuous function
f : X → [0, 1] with f (x0 ) = 1, f (x) = 0 for x ∈ E.
Urysohn [12] introduced these spaces in 1925. They are also called Tychonoff
spaces because of Tychonoff’s work [13]; e.g. see 8.32. A completely regular space
is regular (U = f −1 [0, 1/2) and V = f −1 (1/2, 1] are disjoint open sets containing
E, x0 , respectively,), but there are regular spaces on which the only continuous real
functions are constants (Hewitt [14]). A normal space is completely regular. So,
completely regular spaces lie between T3 spaces and T4 spaces (and are therefore
called T3 21 spaces!).
Normality is the only the separation property considered in the chapter that has
bad behaviour vis à vis subspaces and products.
Definition 8.30 Let X be a space and let F be a family of continuous maps f from
X to a space Y f . F separates points of X if for x = y in X , there is an f ∈ F
such that f (x) = f (y). F separates points and closed sets if for each closed set E
in X and x ∈ E c , there is an f ∈ F such that f (x) ∈
/ f (E). The evaluation map
e : X → f ∈F Y f is defined by e(x)( f ) = f (x).
The next lemma gives the basic properties of the evaluation map. It will also be used
in the proof of the Urysohn metrisation theorem in the next chapter.
Lemma 8.31 (Embedding lemma)
Let notation be as above. Then
1) The evaluation map e is continuous.
2) e is one-one if F separates points of X .
3) e is an open map onto e(X ) if F separates points and closed sets.
Proof One part is easy in view of Tychonoff’s product theorem. Any product of
copies of [0,1] is a compact Hausdorff space, hence is normal. In particular, it is
completely regular and so is any subspace.
Conversely, suppose X is a completely regular space. For each continuous
f : X → [0, 1], let I f = [0, 1] and Y = f I f . (So Y = [0, 1]C(X,[0,1]) .) Take Y f =
[0, 1] for all f ∈ F = C(X, [0, 1]) in the lemma (8.31) to show that the evalua-
tion map given by e(x) = ( f (x)) is a homeomorphism of X onto e(X ). Since X is
completely regular, F separates points and closed sets. This also gives separation of
points as X is T1 . The proof is complete on invoking the lemma (8.31).
(0,1) is not compact but is a dense subspace of the compact space [0,1]. For what
spaces X is such a situation possible?
Example 8.36 [0,1] is a compactification of (0,1), but it is not the Stone-Čech com-
pactification as the continuous function ϕ(x) = sin x1 on (0,1) taking values in [−1,1]
has no continuous extension to [0,1].
In general, β X is huge. For example, βN has the cardinality of the set of all subsets
of R (See [10] or [18]). For more on β X , see [19].
Exercise 8
1. T0 -spaces. Recall the Sierpiǹski space S = {0, 1} with the only proper open
set {1}. Thus {1} is an open set containing 1 but not 0. However there is no open
set containing 0 but not 1, as the only open set containing 0 is the whole space.
Such spaces are called T0 spaces. Thus, X is a T0 space if, for every pair x, y of
distinct points, either there is a neighbourhood of x not containing y or there is a
neighbourhood of y not containing x (but not necessarily both). (If both hold, the
space is a T1 space mentioned earlier.) These spaces are hardly interesting and we
will have no further discussion about them.
Find another T0 space.
2. T1 -spaces. a) The following are equivalent: i) X is a T1 space; ii) one-point
sets are closed in X ; iii) E = ∩{V : E ⊂ V, V is open} for any E ⊂ X ; iv) id X :
(X, T ) → (X, Tco f ) is continuous.
b) In a T1 space, if x is a limit point of E, every neighbourhood of x has infinitely
many points of E. So, finite sets have no limit points.
c) A T1 -space X is countably compact if and only if it is Bolzano–Weierstrass
compact.
d) On any set X , there is a smallest T1 topology.
e) The derived set E of any set in a T1 space is closed.
3. Hausdorff spaces. a) X is Hausdorff if {(x, x ) : f (x) = f (x )} is closed
whenever f : X → Y is an open surjective map.
b) If T is a Hausdorff topology on X , any finer topology is also Hausdorff. What
are the corresponding statements for other properties like compactness and other
separation properties?
c) Every infinite Hausdorff space contains a countably infinite discrete subspace.
8.4 Completely Regular Spaces 121
d) There are normal spaces in which every closed set is a G δ ; such spaces are
called perfectly normal spaces and in such a space every subspace is normal. Metric
spaces are perfectly normal.
e) Every Fσ set in a normal space is normal.
f) If X is normal and E 1 , . . . , E n are closed sets with ∩n1 F j = ∅. Then there are
open sets V j ⊃ E j with ∩n1 V̄ j = ∅.
8. Paracompact spaces. A cover U of a space X is called a refinement of a
cover V if every U ∈ U is a subset of some V ∈ V .
a) X is compact if and only if every open cover of X has a finite refinement that
covers X .
A space X is called a paracompact space if every open cover of X has a locally
finite refinement that is a cover.
These spaces were introduced by Dieudonné in 1944 [23]. They are very useful
in topology and are important in geometry because of the existence of partitions of
unity. A.H.Stone is one of the most important contributors to the study of paracom-
pactness in topology.
b) Every paracompact Hausdorff space is normal. (Theorem of A.H.Stone.) ( See,
e.g., Dugundji [10], for the results given here.)
c) Every metric space is paracompact.
9. Quotients. a) Let R be an equivalence relation on a space X and let p : X →
X/R be the canonical map.
i) If p is open and R is closed in X × X , then X/R is Hausdorff.
ii) X/R is T1 if and only if each equivalence class is closed in X .
iii) Suppose p is open and closed. If X is normal, then so is X/R.
b) Let X, Y be compact Hausdorff spaces and let f : X → Y be a continuous
surjection. Let R be the equivalence relation whose classes are the fibres of f , i.e.
x Ry if f (x) = f (y). Then the quotient space X/R is homeomorphic to Y . In fact,
h which takes the fibre f −1 (x) to f (x) is well defined and is a homeomorphism. We
have seen examples of this in Chap. 6: S1 as a quotient of [0,1] and the torus as a
quotient of the square.
10. Cantor set. Let C be the Cantor ternary set in [0,1]. Every x ∈ C has a ternary
(or triadic) expansion x = a3nn , where (an ) ∈ {0, 2}N .
a) ϕ : {0, 2}N → C, (an ) → a3nn , is a bijection.
b) If C has the subspace topology from R and {0, 2}N is given the product topology,
ϕ is continuous.
c) ϕ is a homeomorphism.
d) Construct a continuous function on [0,1] whose zero set is C.
11. Space filling curves: existence. a) ψ : {0, 2}N → [0, 1] defined by
ψ((an )) = 2an+1
n
is a continuous surjection.
b) There is a continuous map of the Cantor set C onto [0,1].
c) C is homeomorphic to C × C.
d) There is a continuous surjection of C onto [0, 1] × [0, 1].
8.4 Completely Regular Spaces 123
8.5.1 Tietze
Heinrich Tietze (1880–1964) was an Austrian mathematician, best known for his
work in topology. He was the one who defined normal spaces [4], 1923 and is famous
for the Tietze Extension Theorem. He is considered as one of the first ‘topology
specialists’ and carried forward Poincaré’s work on topology. In a long 1908 paper,
he settled many outstanding questions. For example, he produced a finite presentation
for the fundamental group and invented the now well-known Tietze transformations to
show that fundamental groups are topological invariants. He gave the first reference to
the isomorphism problem for groups: if two groups are defined by finite presentations,
is there an algorithm to decide whether they are isomorphic or not? He also completed
the relationship between the fundamental group and what is now called the first
homology group. For more details, see the account of I.M.James [25]. He also worked
124 8 Separation Axioms
on other areas like geometry, knot theory, map colouring, continued fractions, prime
numbers, ruler-compass constructions, partitions.
8.5.2 Urysohn
8.5.3 Carathéodory
8.5.4 M. H. Stone
Marshall Stone (1903–1989), was an American mathematician, best known for his
work on Functional Analysis. His thesis adviser at Harvard was George D.Birkhoff,
a leading mathematician at that time. He is also well-known for rebuilding the Math-
ematics Department at Chicago into one of eminence again after World War II. Sev-
eral analysts of repute in the twentieth century were among his students. Stone-Čech
compactification that we have discussed, Stone–Weierstrass Theorem generalising
the classical Weierstrass Approximation Theorem, Stone–von Neumann theorem on
the canonical commutation relations in Quantum Mechanics/ representations of the
Heisenberg group, Stone’s theorem on one-parameter groups of unitary operators,
Banach–Stone theorem and Stone’s representation theorem for Boolean algebras are
all famous (and all of them were discovered in the 1930s!). His monograph Linear
transformations on Hilbert spaces (1932, incidentally the same year as Banach’s
classic Théorie des opérations linéaires appeared) was one of the earliest on the
topic and is considered a classic. Of interest in India is the fact that he visited India
frequently and died in Madras (now called Chennai).
References
18. Wilder, R.L.: Evolution of the topological concept of “connected". Am. Math. Monthly 85,
720–726 (1978). https://doi.org/10.1080/00029890.1978.11994684, Correction and Adden-
dum, Am. Math. Monthly 87 31–32 (1980). https://doi.org/10.1080/00029890.1980.11994947
19. Gillman, L., Jerison, M.: Rings of Continuous Functions. Van Nostrand (1960)
20. Munkres, J.R.: Topology, 2nd edn. Pearson Education, Asia (2001)
21. Steen, L.A., Seebach Jr., J.A.: Counter Examples in Topology. Dover (1995)
22. Engelking, R.: General Topology Revised Edition. Heldermann Verlag (1989)
23. Dieudonné, J.A.: Une Généralisations des Espaces Compacts. J. Math. Pures. Appl. 23, 65–76
(1944)
24. Peano, G.: Sur une Courbe qui ramplit Toute une Plane. Math. Ann. 36, 157–160 (1890). https://
doi.org/10.1007/BF01199438
25. James, I.M.: From combinatorial topology to algebraic topology. In: James, I.M. (ed.) History
of Topology, pp. 561–574. Elsevier (1999)
Chapter 9
Connected Spaces
A satisfactory formulation for the concept of connectedness was obtained only after
many inadequate starts. The evolution took a long time—through Bolzano (1817
[1]), Cantor (1883 [2]), C.Jordan (1893 [3]), Schönfliesz (1904 [4]) and others—
and culminated in the modern formulation in 1906, due to F.Riesz and N.J.Lennes,
independently.
‘Das mathematische Kontinuum heisse zusammenhangende wenn es nicht in zwei
offene Teilmengen zerlegt werden kann, die Komplementarmengen fur einander sein.’
(F. Riesz, 1906, [5])
(‘The mathematical continuum is called connected if it cannot be decomposed into
two complementary open subsets.’)
‘A set of points is a “connected set" if at least one of any two complementary
subsets contains a limit-point of points in the other set.’ (Lennes, 1906, [6])
Riesz’s formulation says that the space cannot be decomposed into two com-
plementary open sets and is equivalent to that of Lennes. It was rediscovered by
Hausdorff in 1914. For the history of evolution of the concept of connectedness, see
[7].
We begin with an important, strong, geometric form of connectedness before
moving on to a more abstract form and then on to local versions of these properties.
An interval, the real line, the graph of a continuous real function on the real line,
the punctured plane, an annulus, a circle, the surface of a sphere—all these share an
important geometric property: any two points in any of these spaces can be connected
by a path which lies entirely in the space. We now take up this geometric form of
connectedness. This concept is of considerable importance in geometric aspects of
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 127
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_9
128 9 Connected Spaces
topology and is more intuition friendly than the more abstract property to be taken
up in the next section.
Definition 9.1 A path in a topological space X is a continuous map γ : [0, 1] → X .
For such a path, x0 = γ (0) is called the initial point and x1 = γ (1) is called the
terminal point of γ . In this case we also say that γ is a path from x0 to x1 or a path
joining x0 and x1 . The space X is said to be path connected if, for every pair x, y of
points in X , there is a path in X from x to y. A path connected set in X is a subset
C that is path connected in the relative topology.
The concept of paths and path connected sets have been around for a long time,
going back at least to the time of Weierstrass. Here are some simple properties of
path connected spaces.
Proof Let X be path connected and let f : X → Y be continuous and surjective. Let
y0 = f (x0 ), y1 = f (x1 ) ∈ Y . If γ is a path in X from x0 to x1 , then then γ f := f ◦ γ
is a path in Y from y0 to y1 .
Dealing with products is easy for path connected spaces, in contrast to products
of connected spaces considered in the next section.
Proposition
9.3 If {X i } is a collection of path connected spaces, the product space
X = X i is also path connected.
Proof Let x = (xi ), y = (yi ) ∈ X and let, for each i, γi be a path in X i from xi to
yi in X i . Then γ (t) = (γi (t)), 0 ≤ t ≤ 1. gives a path in X from x to y. (Each X i is
the image of X under the projection pi and so the converse is obvious.)
There is a concept of ‘connectedness’ more general, more abstract and less tangible
than the geometric property of path connectedness discussed in the previous section.
It is an important concept in geometry and analysis. Roughly speaking, these are
spaces that are in ‘one piece’. But to give a mathematical shape to this vague notion
of connectedness, we start with a proposition.
Proposition 9.8 For a space X , the following are equivalent:
1) If A, B are nonempty subsets and if X = A ∪ B, then either A ∩ B̄ or Ā ∩ B
is nonempty.
2) If X = A ∪ B, A, B disjoint open sets, then either A or B is empty.
3) If X = A ∪ B, A, B disjoint closed sets, then either A or B is empty.
4) If A ⊂ X is both open and closed then either A is empty or A = X .
5) If f : X → ({0, 1}, Td ) is continuous, it is a constant.
Example 9.10 1. Intervals in R are connected, as we shall soon see. This may look
obvious, but the proof is actually not all that trivial.
2. A discrete space with more than one point is not connected, whereas any
indiscrete space is connected.
3. An infinite set with cofinite topology is connected (since no two nonempty
open sets are disjoint!).
4. [0,1] ∪ [2,3] is not connected.
132 9 Connected Spaces
2
5. The space O(n) of real n × n orthogonal matrices (as a subspace of Rn ) is
not connected since the determinant function is a continuous function mapping O(n)
onto (check!) the discrete space {1, −1}.
More work is needed before important, nontrivial examples can be given. The fol-
lowing is almost trivial, but is very important.
Exercise. Give a proof of 9.11 using the characterisation (5) of 9.8. Note also that,
conversely, 9.11 implies (5) of 9.8.
As the most basic examples, let us begin by looking at connected subsets of the
real line. In R, intervals are the only connected sets.
Thus the list of connected subsets of R is: {a}; [a, b], [a, b), (a, b], (a, b); (−∞, a],
(−∞, a), [a, ∞), (a, ∞); (−∞, ∞) = R, a, b ∈ R, a < b.
Here is the intermediate value theorem with which the book started.
Corollary 9.13 (Intermediate value theorem)
If f is a continuous, real valued function on an interval in R, then f takes every
value between any two of its values.
{x} × Y {x } × Y
E(x) E(x )
X × {y0 }
(x, y0 ) (x0 , y0 ) (x , y0 )
We can handle the infinite product case utilising the finite product case and the
following lemma,
Proof It suffices to show that D intersects any basic open set B = ∩i∈F pi−1 (Vi ),
where Vi is open in X i and F is a finite subset of I . But this is all but obvious: taking
xi = ai , i ∈
/ F and xi ∈ Vi arbitrary for i ∈ F, we see that x = (xi ) ∈ B ∩ D.
Theorem 9.22 Let X be a topological space. For each x ∈ X , let C x denote the
union of all connected subsets of X containing x. Then
9.2 Connected Spaces 135
Proposition 9.26 If, for every pair x, y of points in a space X , there is a connected set
C = C x,y ⊂ X containing both, then X is connected. In particular, a path connected
space is connected.
The converse is not true. A connected space may not be path connected. Here are
two examples.
Example 9.27 1. Let C be the following subset of the plane: the segment (0,1] on
the horizontal x-axis together with vertical line segments of unit length at each point
( n1 , 0), n = 1, 2, . . . (Fig. 9.5). Thus C is the union (0, 1] × {0} ∪ (∪∞
1 { n } × [0, 1]).
1
Then C is path connected. In fact, any two points in C can be joined by a path in C
consisting of at most three line segments. Hence C̄ = C ∪ {0} × [0, 1] is connected.
However, C̄ is not path connected. For example, there is no path in C̄ from the point
(0,1) to the point (1,0). Thus a path connected space may not have a path connected
closure and a connected space may not be path connected. Observe also that C̄ has a
single connected component and two path components, namely C and the segment
{0} × [0, 1]. C is usually referred to as the comb space.
2. Another example with the same properties is the topologist’s sine curve.
Consider G f = {(x, sin(1/x)) : 0 < x ≤ 1}, the graph of the continuous function
f (x) = sin(1/x), 0 < x ≤ 1 (Fig. 9.6). It is path connected, but it closure Ḡ f =
G f ∪ {0} × [0, 1] is not path connected. Here again there is no path from (0,1) to
(1,0) in Ḡ f .
9.2 Connected Spaces 137
sin(1/x)
However, we have the following positive result; it is useful, for example, in com-
plex analysis (e.g. Cauchy’s integral theorem).
Exercise. 1. Any two points in an open connected set C in Rn can actually be con-
nected by a polygonal arc in C. For example, in the plane there is such a polygonal
path made up of horizontal and vertical line segments.
2. Any product of discrete spaces is totally disconnected.
3. For x ∈ X, y ∈ Y , is the connected component of (x, y) in X × Y the product
C x × C y of the connected components of x and y?
Local analogues of the properties discussed in the last two sections are important in
geometric considerations in topology and analysis.
Locally connected spaces were introduced by Hahn (1914), [10], who proved 9.30
for metric spaces.
Note that discrete spaces (with more than one point) are not connected but are
locally connected. The comb space C and topologist’s sine curve G f are locally
connected, but their closures are not locally connected; for instance, the defining
property fails for the point (0,1).
138 9 Connected Spaces
9.4.1 Jordan
Camille Jordan (1838–1922) was a French mathematician well-known for his work
on group theory and analysis. Jordan curve theorem (a fundamental theorem in plane
topology), Jordan measure, Jordan decomposition, Jordan matrix, Jordan canonical
form for matrices, Jordan–Höder theorem for groups are important contributions
from him. His Cours d’Analyse was a popular analysis text book in the nineteenth
century. His work in group theory helped to bring Galois theory (discovered by the
French teenager genius Evariste Galois) to the mainstream of mathematics.
9.4.2 Hahn
Hans Hahn (1879–1934) was a lawyer turned mathematician from Austria, who was
also a philosopher. He is best known for his work in functional analysis. His most
famous contributions include the Hahn–Banach theorem (a fundamental theorem
of functional analysis) and the Hahn decomposition theorem in measure theory.
Vitali–Hahn–Saks theorem and Hahn–Mazurkiewicz theorem are also well-known.
He made important contributions to real analysis, calculus of variations and topology
(the concept of local connectedness is due to him). His famous students include Karl
Menger (dimension theory), Witold Hurewicz (higher homotopy groups, dimension
theory) and Kurt Gödel (set theory and logic, the continuum hypothesis, the famous
Gödel incompleteness theorem).
9.4.3 Kuratowski
removing a single point! His monographs on topology are famous. During World War
2, he lectured at the ‘underground university’ and after the war played an important
role in the revival of mathematics and science in Poland.
9.4.4 Knaster
Bronisław Knaster (1893–1980) of Poland is known for his work in Set Topology.
He was a student of Mazurkiewicz. His hereditarily indecomposable continuum,
the Knaster continuum, Knaster–Kuratowski fan (mentioned above), Knaster–Tarski
theorem on lattices, his work with Banach and Steinhaus on ‘fair division’ are well-
known. He was one of the trio to prove Brouwer’s theorem from Sperner’s lemma. It
is interesting to note that he took to mathematics when he was past 20; he had studied
medicine for three years at Paris earlier. He had a passion for mathematical editorial
work and took up editorial work for the Polish mathematical journals Studia Math-
ematica and Colloquium Mathematicum. He was the one who translated Banach’s
celebrated monograph Théorie des opérations linéaires into French.
References
1. Bolzano, B.: Rein analytischer Beweis des Lehrsatzes, dass zwischen je zwei Werthen, die
ein entgegengesetztes Resultat gewähren, wenigstens einer reelle Wurzel der Gleichung liege.
Prague, Gottlieb Haase (1817)
2. Cantor, G.: Über unendliche, lineare Punktmannigfaltigkeiten, Math. Ann. 15 1–7 (1879);
17, 355–388 (1880); 20, 113–121 (1882); 21 51–58 (1883), 545–591; 23, 453–488
(1884). https://doi.org/10.1007/BF01444101, https://doi.org/10.1007/BF01446232. https://
doi.org/10.1007/BF01443330, https://doi.org/10.1007/BF01442612. https://doi.org/10.1007/
BF01446819, https://doi.org/10.1007/BF01446598
3. Jordan, C.: Cours d’Analyse de l’École Polytechnique, vol. I., 2nd edn. Paris (1893)
4. Schoenflies, A.: Beiträge zur Theorie der Punktmengen. Math. Ann. 58, 195–234 (1903).
https://doi.org/10.1007/BF01447784
5. Riesz, F.: Die Genesis des Raumbegriffs. Math. Naturwiss. Ber. Ungarn. 24, 309–353 (1906)
6. Lennes, N.J.: Curves in non-metrical analysis situs with an application in the calculus of
variations. Am. J. Math. 33, 287–326 (1911)
7. Wilder, R.L.: Evolution of the topological concept of “connected". Am. Math. Monthly 85,
720–726 (1978). https://doi.org/10.1080/00029890.1978.11994684, Correction and Adden-
dum, Am. Math. Monthly 87 31–32 (1980). https://doi.org/10.1080/00029890.1980.11994947
8. Knaster, B., Kuratowski, K.: Sur les ensembles connexes. Fund. Math. 2, 206–255 (1921).
https://doi.org/10.4064/fm-2-1-206-255
9. Hausdorff, F. Grundzüge der Mengenlehre. Leipzig (1914)
10. Hahn, H.: Über der allgemeinste eben Punktmenge, die stetiges Bild einer Strecke ist Jahresber.
Deut. Math. Ver. 23, 318–322 (1914). https://doi.org/10.1007/978-3-7091-6601-711
11. Tietze, H.: Beiträge zur allgemeinen Topologie III, Monatsh.f. Math. und. Phys. 32, 15–17
(1923). https://doi.org/10.1007/BF01705586
12. Dugundji, J.: Topology. Prentice Hall of India (1975)
13. Kelley, J.L.: General Topology. Affiliated East-West Press (1969)
14. James, I.M.: From combinatorial topology to algebraic topology. In: James, I.M. (ed.) History
of Topology, pp. 561–574. Elsevier (1999)
Chapter 10
Countability Axioms
Most of us are more comfortable dealing with countable sets than with uncountable
sets. In this chapter we deal with various properties of topological spaces that involve
countable sets in some way. Hausdorff, in his Grundzüge [1] (1914), gives two
countability axioms (which are now called the first and second countability axioms.):
Abzählbarkeitsaxiome: (E) Für jeder Punkt x ist die Menge seiner verschiedenen
Umgebungen Ux höchsten abzählbar.
(F) Die Menge aller verschiedenen Umgebungen U ist abzählbar.
(Countability axioms: (E): For each point x, the set of all different neighbourhoods
Ux is countable.
‘(F): The set of different neighbourhoods U is countable’.)
Fréchet [2] defined another countability condition thus:
“Nous appellerons ensuite classe séparable une classe qui puisse être considérée
d’au moins une façon comme l’ensemble dérivé d’un ensemble dénombrable de ses
propres éléments.” (Fréchet, [2], p. 23) (“We will call a class separable if it can be
considered in at least one way as the derived set of a countable set of its elements.”)
Proof 1) Let x be a limit point of A and let {Un } be a local base at x. Choose
xn ∈ Un ∩ A, xn = x for each n. Then xn → x.
2) It suffices to show that f ( Ā) ⊂ f (A) for any A ⊂ X . If x ∈ Ā, by (i), choose
xn ∈ A, xn → x. Then f (xn ) → f (x), so f (x) ∈ f (A).
Definition 10.4 A topological space X having a countable base for its topology is
called a second countable space.
Example 10.5 1. R is second countable: open intervals (a, b), with rational a, b,
form a countable base.
2. Every second countable space X is first countable.
3. R with the discrete topology is first, but not second, countable.
4. R is not second countable (but is first countable, as already noted). Let B be
any base for R . For any x, the interval [x, x + 1) is an open set and so there is a
Bx ∈ B such that x ∈ Bx ⊂ [x, x + 1). Since Bx = B y for x = y (a is min Ba for
every a), the collection {Bx : x ∈ R } is uncountable. Thus B is uncountable. We
have shown that any base for R is necessarily uncountable. We can also conclude
that R × R is not second countable from the fact that it contains an uncountable
discrete subspace (see 8.16).
Proof Let X be a space with a countable base {Bn } and let {Ui }i∈I be an open cover
for X . Let J = {n : Bn ⊂ Ui f or some i ∈ I }. For each n ∈ J , let i n ∈ I be such
that Bn ⊂ Uin . If x ∈ X , there is a Ui with x ∈ Ui and then there is a k such that
x ∈ Bk ⊂ Ui . This means that k ∈ J and x ∈ Uik . In other words, {Uin : n ∈ J } is a
cover for X and is a countable subcover that we are looking for.
10.1 Countability Properties 145
In view of this theorem [3], a space for which every open cover has a countable
subcover is called a Lindelöf space. Ernst Lindelöf was a mathematician from Finland
and Lars Ahlfors’ teacher.
Example 10.8 1. Every compact space is a Lindelöf space. R is an example of a
Lindelöf space that is noncompact.
2. R is Lindelöf.
Definition 10.9 A topological space is said to be separable if it has a countable
dense subset. Fréchet introduced this property.
Proposition 10.10 A second countable space is separable.
Proof Suppose {Bn } is a countable base for a space X . For each n, let xn ∈ Bn . The
countable set D of all the xn is dense in X . For, every nonempty open set U contains
a Bn and so contains the point xn . Thus D intersects every nonempty open set and
hence is dense.
Example 10.11 R is separable (Q is dense in R !), but is not second countable.
Thus R satisfies all the countability properties considered in this chapter except
second countability.
In a metric space, separability is equivalent to second countability.
Theorem 10.12 For a metric space X , the following are equivalent:
1) X is second countable.
2) X is Lindelöf.
3) X is separable.
Proof We have seen that 1) implies both 2) and 3) for any space X . For a metric
space X , we show 2)⇒ 3)⇒ 1).
2) ⇒ 3): For each n, the balls B(x, 1/n), x ∈ X, form an open cover Bn of X , so
has a countable subcover Bmn = {Bmn = B(xmn , 1/n)}m . The countable set {xmn } is
dense in X .
3) ⇒ 1): Let D = {xn } be a countable dense set in a metric space X . We prove
that the countable collection {B(xm , 1/n) : m, n ∈ N} of open balls form a base
for X . If U is open and x ∈ U , then B(x, 1/n) ⊂ U for some n and then there
is an xm ∈ B(x, 1/2n) ∩ D, as D is dense. So x ∈ B(xm , 1/2n) ⊂ B(x, 1/n) ⊂ U ,
completing the proof.
question arises whether ... regularity alone is not enough for the metrisability of
spaces satisfying the second axiom of countability”.) Tychonoff answered this affir-
matively in [6] by proving that a second countable regular space is, in fact, normal.
Tychonoff was 19 at that time!
ii) The general metrisation theorem, giving necessary and sufficient conditions
for metrisation of a topological space, was obtained, independently, by three math-
ematicians in three different countries: Bing (USA), Nagata (Japan) and Smirnov
(Soviet Union) [7–9]. See [10–13] for details.
Exercise 10
10.3.1 Lindelöf
hypothesis on the rate of growth of the Riemann zeta function on the critical line are
some of his significant contributions. Lars Ahlfors, the great complex analyst who
was one of the first two Fields medallists (1936), was his student at Helsinki.
References
Not all important spaces are compact—even the real line is not. Locally compact
spaces, introduced, independently, by Tietze [1] and Alexandroff [2], are nice enough
and sufficiently abundant to be very important in analysis and geometry.
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 151
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_11
152 11 Locally Compact Spaces
Sets with compact closure are called relatively compact or precompact sets. Thus X is
locally compact means that every point of X has a precompact open neighbourhood.
Example 11.2 Any compact space is locally compact and so is any discrete space.
The Euclidean space Rn and an infinite discrete space are locally compact and non-
compact. Q is not locally compact.
Theorem 11.3 In a locally compact Hausdorff space X , every neighbourhood of
any x ∈ X contains a compact neighbourhood.
Proof Let x ∈ X and let V be an open neighbourhood of x with V̄ compact. Then
V̄ is a compact Hausdorff space, hence is normal and is therefore regular. If U is
a neighbourhood of x, U ∩ V̄ is a neighbourhood of x in this regular space and so
there is a relative neighbourhood E of x with U ∩ V̄ ⊃ Ē V̄ := the closure of E
in V̄ = Ē ∩ V̄ . But E = V̄ ∩ W , with W an open neighbourhood of x in X . Now
V ∩ W is a neighbourhood of x in X with compact closure, being a closed subset of
V̄ and is contained in U .
This property extends to compact sets in place of points: every neighbourhood of
a compact set contains a compact neighbourhood.
Corollary 11.4 Let X be a locally compact Hausdorff space, K be a compact subset
and U be an open subset such that K ⊂ U . Then there is a precompact open set V
such that K ⊂ V ⊂ V̄ ⊂ U .
Proof For each x ∈ K there is a precompact open set Vx such that x ∈ Vx ⊂ V̄x ⊂ U .
By compactness of K , the open cover {Vx : x ∈ K } of K has a finite cover {Vx j }m 1 . If
V = ∪m 1 Vx j
, then V̄ = ∪m
1 V̄x j
and so V̄ , being a finite union of compact sets, is com-
pact. Thus V is a precompact set with K ⊂ V ⊂ V̄ ⊂ U , and the result is proved.
Corollary 11.5 A locally compact Hausdorff space has a basis of open precompact
sets.
Proof This is immediate from the theorem.
If the space is also second countable, we can improve on this.
Proposition 11.6 A locally compact Hausdorff space X with a countable basis has
a countable base consisting of precompact sets.
Proof Let {Bn } be a countable basis for X . Cover each Bn by precompact open sets,
{Wx : x ∈ Bn }. As X is second countable, so is Bn , hence is a Lindelöf space. So the
open cover {Wx : x ∈ Bn } has a countable subcover, say {Wn,k }k∈N . The countable
collection {Wn,k : k, n ∈ N} of open precompact sets form a basis for X .
Proposition 11.7 A subspace E of a locally compact Hausdorff space X is locally
compact if and only if E is locally closed in X , or equivalently, E = V ∩ F where
V is open and F is closed in X . In particular, open sets and closed sets in X are
locally compact.
11.1 Local Compactness 153
Example 11.8 The space M(n, R) of matrices, with the usual Euclidean topology, is
locally compact and Hausdorff. Hence so is the open subset G L(n, R) of nonsingular
matrices. The space S L(n, R) of matrices of determinant one is closed and so is
locally compact too.
The intersection of two dense sets may not be dense (it can even be empty!), but
if the sets are also open, the intersection (of even countably infinitely many of them)
is dense. The next two results are valid also in complete metric spaces; see 12.14 and
12.15.
Theorem 11.9 (Baire’s theorem)
If {Vn }∞
1 is a countable collection of open, dense subsets in a locally compact Haus-
dorff space X , then ∩Vn is also dense.
A set E whose closure has empty interior is called a nowhere dense set. Note that E
is nowhere dense if and only if E c is dense. For remarks on the term ‘category’ see
remarks after 12.15
Example 11.11 Consider R and let Vq = {q}c , q ∈ Q. Each Vq is open and dense
and ∩q Vq is the set of irrational numbers, which is dense. This also shows that the
set of irrational numbers is a G δ . Now Baire’s theorem implies that Q is not a G δ .
For, if Q = ∩Un is a countable intersection of open sets Un , then Un ⊃ Q for each
154 11 Locally Compact Spaces
n and so Un is dense. But then {Vq } ∪ {Un } is a countable family of open dense sets
in R with empty intersection, contradicting Baire’s theorem. The same arguments
show that no countable dense set is a G δ in R.
In the presence of compactness, we have seen that Hausdorff spaces are normal. If
only local compactness is assumed, we can still get complete regularity of Hausdorff
spaces.
Proof We have already seen that products of Hausdorff spaces are Hausdorff, so
we need to show that X × Y is locally compact. This is easy. Let x ∈ X, y ∈ Y .
Then there are open sets U ⊂ X, V ⊂ Y with Ū and V̄ compact. Then U × V is an
open neighbourhood of (x, y) in X × Y with compact closure Ū × V̄ . The proof is
complete.
Remark 11.16 In the converse part, from the local compactness of X we can actually
conclude that the X i are locally compact from the facts that the projections are open
surjections.
P*
P
Q*
Fig. 11.2 Sn , the one-point compactification of Rn via stereographic projection: N = north pole,
P, Q ∈ Rn ,P ∗ , Q ∗ images on Sn
(a,1) (b,1)
Exercise 11
11.3.1 Alexandroff
11.3.2 Dieudonné
Jean Dieudonné (1906–1992) was a French mathematician and was one of the
founders of the Bourbaki group. His thesis, as a student of Paul Montel, was on
classical analysis. He worked in a wide variety of mathematical areas including gen-
eral topology, topological vector spaces, algebraic geometry, invariant theory and the
classical groups. He was the scribe for Bourbaki: Armand Borel mentions ‘the super-
human efficiency of Dieudonné’ and says ‘he took care of the final drafts, exercises,
and preparation for the printer of all the volumes’ [6]. (Quoted with the permission of
11.3 Biographical Notes 161
References
1. Tietze, H.: Beiträge zur allgemeinen Topologie II. Über die Einführung uneigentlicher Elemente.
Math. Ann. 91, 210–224 (1924). https://doi.org/10.1007/BF01556079
2. Alexandroff, P.: Über die Metrisation der im Kleinen kompakten topologische Räume. Math.
Ann. 92, 294–301 (1924). https://doi.org/10.1007/BF01448011
3. Bourbaki, N.: General Topology, Chapters 1–4. Springer (1995)
4. Engelking, R.: General Topology Revised Edition. Heldermann Verlag (1989)
5. Kim, S.S.: A characterization of the set of points of continuity of a real function. Am. Math.
Monthly 106, 258–259 (1999). https://doi.org/10.1080/00029890.1999.12005038
6. Borel, A.: Twenty-Five Years with Nicolas Bourbaki (1949–1973). Not. Am. Math. Soc. 45,
373–380 (1998)
Chapter 12
Complete Metric Spaces
Proof If the terms of a sequence are eventually close to some point, they are also
eventually close to each other. The reader should write out a formal proof (which
takes hardly more space).
So, the reverse implication is the one of interest and importance. Recall that the
property of R on which most of the basic results of real analysis is based is its
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 163
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_12
164 12 Complete Metric Spaces
completeness. In fact, the chief reason for the construction of real numbers from
rational numbers is the incompleteness of Q and R was obtained by ‘completing’ Q.
Example 12.3 The basic fact on which completeness of function spaces and sequence
spaces rests is the completeness of R.
1. Rn and Cn are complete for all n with each of the metrics d1 , d2 and d∞ (which
are defined on Cn just as for Rn ).
2. For any metric space X , the space Cb (X ) of bounded (real or complex) continu-
ous functions on X is complete with the sup metric: dsup ( f, g) =
supx∈X | f (x) − g(x) | is complete. (The basic underlying reason: ‘uniform limits
of continuous functions are continuous’.)
3. The sequence spaces 1 , 2 and ∞ given earlier are all complete.
4. A discrete metric space is complete: in a discrete metric space, a Cauchy
sequence is eventually a constant and hence is convergent.
1 ∞
5. In (0,1), { n+1 }1 is a Cauchy sequence having no limit.
Which subspaces of complete spaces are complete?
Proposition 12.4 Let X be a metric space and let E ⊂ X .
1) If E is complete (with the induced metric), then it is closed in X (i.e. a complete
subspace of a metric space is closed.)
2) If X is complete and E is closed in X , then E is complete (i.e. a closed subspace
of a complete metric space is complete).
Our discussion shows that any uniformly convergent sequence of continuous func-
tions on [0,1] yields an equicontinuous family. The next result, important in many
applications, is in the reverse direction. It is a simple modern version of results first
proved by Ascoli [6] in 1883 and Arzelà [1] in 1895. It is used in many parts of anal-
ysis, e.g. normal families in complex analysis. Our formulation is slightly different
from the usual one.
Theorem 12.10 (Ascoli–Arzelà theorem)
Let X be a separable metric space and let Y be a complete metric space having
the Bolzano–Weierstrass property. If F is an equicontinuous, point-wise bounded
family (i.e. the set { f (x) : f ∈ F } is bounded in Y for each x ∈ X ) of maps from
X into Y, then every sequence in F has a subsequence converging uniformly on
compact subsets of X .
Proof Let {xn } be a countable dense set in X and let { f n } be a sequence in F . Then
{ f n (x1 )} is a bounded sequence in Y and so, by the assumption on Y , there is a
subsequence { f n1 (x1 )} which converges in Y . Next consider the bounded sequence
{ f n1 (x2 )} and get a convergent subsequence { f n2 (x2 )}. Note that both { f n2 (x1 )} and
{ f n2 (x2 )} converge. Proceed inductively to get subsequences { f nk } with { f nk (x j )}
convergent for each 1 ≤ j ≤ k such that each of them is a subsequence of the pre-
ceding one. The ‘diagonal’ subsequence {gn = f nn } has the property that {gn (xk )}n
converges for all k. (Cantor’s diagonal argument again!)
Let K ⊂ X be compact. Now let us invoke the equicontinuity. Let ε > 0 and
choose δ > 0 such that d X (x, y) < δ implies dY ( f n (x), f n (y)) < ε/3 for all n.
Compactness of K ensures that it can be covered by a finite number of balls
12.1 Completeness and Ascoli–Arzelà 167
f (x) − g(x)
. Then any equicontinuous, point-wise bounded family F is precom-
pact in C(X, Rn ).
Proof If F is equicontinuous, point-wise bounded, then so is its closure F¯ in
C(X, Rn ). It is sequentially compact by the theorem.
Baire’s theorem for complete metric spaces is an important tool to obtain some basic
theorems of Functional Analysis and has other applications as well. Banach’s fixed
point theorem is used in differential equations and other parts of analysis. We take
a rather unusual route to Baire’s theorem via the following result of Bourbaki [2].
Theorem 12.13 (Bourbaki’s Mittag-Leffler Theorem)
Let (X n , dn ) be a complete metric
space for each n and f n : X n+1 → X n be con-
tinuous maps. Let X = {(xn ) ∈ n X n : xn = f n (xn+1 )}. If the range f n (X n+1 ) is
dense in X n for each n, then p1 (X ) is dense in X 1 , where p1 is the projection on
X 1.
Proof First let us define metrics d̃n on the X n inductively as follows. Let d̃1 =
d1 and d̃n+1 (x, y) = dn+1 (x, y) + d̃n ( f n (x), f n (y)) for x, y ∈ X n+1 , n ≥ 1. Observe
that dn+1 (x, y) ≤ d̃n+1 (x, y) and d̃n ( f n (x), f n (y)) ≤ d̃n+1 (x, y). Moreover,
i) d̃n gives the same topology as dn because dn (xk , x) → 0 if and only if
d̃n (xk , x) → 0;
168 12 Complete Metric Spaces
ii) d̃n is complete; for if {xk } is d̃n -Cauchy, then it is dn -Cauchy, hence by dn -
completeness, dn (xk , x) → 0 for some x ∈ X n and so d̃n (xk , x) → 0.
Let x1 ∈ X 1 and let ε > 0. Since f 1 (X 2 ) is dense in X 1 and f 2 (X 3 ) is dense in X 2
there are x2 ∈ X 2 such that d̃1 (x1 , f 1 (x2 )) < ε/2 and x3 ∈ X 3 with d̃2 (x2 , f 2 (x3 )) <
ε/22 . Then d̃1 ( f 1 (x2 ), f 1 ( f 2 (x3 )) < ε/22 as well. Inductively, choose {xn } with the
following properties:
a) dn (xn , f n (xn+1 )) < ε/2n and
b) d̃k ( f k,n−1 (xn ), f kn (xn+1 )) < ε/2n , n ≥ 2, k < n, where f kn is the composition
fk ◦ · · · ◦ fn .
For each k, { f k,n−1 (xn )}n≥k is a Cauchy sequence in X k and so has a limit ξk . Since
f k,n−1 (xn ) = f k ( f k+1,n−1 (xn )). continuity of f k yields ξk = f k (ξk+1 ), so (ξk ) ∈ X .
We complete the proof by showing that d̃1 (x1 , ξ1 ) ≤ ε by the following estimates:
A system (X n , f n ) as in the theorem is called a projective system or an inverse
system, and X is called the projective limit or the inverse limit of the system. The
theorem (in a much more general form) is given in Bourbaki’s tome [2]. Bourbaki
calls it Mittag-Leffler’s theorem and derives the classical Mittag-Leffler’s theorem on
meromorphic functions from it. Our presentation follows those in Esterle [7] (where
it is used in the theory of Banach algebras) and Runde [8].
Corollary 12.14 (Baire’s theorem, [3])
If {Un } is a sequence of open, dense subsets in a complete metric space X , the n Un
is also dense in X .
Proof Let U, V be open dense sets. Then for any nonempty open set W , V ∩ W
is nonempty as V is dense. It is also open and denseness of U implies W ∩ U ∩ V
is nonempty. Thus U ∩ V intersects each nonempty open set and hence is dense.
Induction gives the result that any finite intersection ∩n1 Uk of open dense sets is
dense.
Consider now the general case. Let {Un } be a sequence of open, dense subsets in
X andlet Vn = ∩1 Uk . Then {Vn } is a decreasing sequence of dense ∞
n
open sets and
U = Vn = Un . We apply the theorem to the sequence {(X n , dn )}0 of complete
spaces, where (X 0 , d0 ) = (X, d) and (X n , dn ) = (Vn , dVn ), n > 0, with dVn as in
12.5, and with the maps f n as the inclusion maps ιn : X n → X n−1 . Clearly the
condition of the theorem is satisfied,
X = {(xn ) ∈ X n : xn = xn−1 , n ≥1} and
p0 (X ) = {x0 ∈ X 0 : x0 ∈ Vn ∀n} = Vn .
The proof is complete.
12.2 Bourbaki, Baire and Banach 169
12.3 Completion
i
X X̄
j σ
X
enough, then y = f¯(x), y = f¯(x ) are close and thus f¯ is uniformly continuous.
The reader can make these precise with ε, δ etc.
Finally, suppose f is an isometry. Then, with notation as before,
dY ( f¯(x), f¯(x )) = lim dY ( f (xn ), f (xn )) = lim d X (xn , xn ) = d X (x, x ).
Corollary 12.21 A uniformly continuous map f on a dense subspace of a metric
space X into a complete metric space has a unique continuous extension to X and
the extension is an isometry if f is.
Now it is easy to get the uniqueness of completion.
Theorem 12.22 Let (X, d) be a metric space. If i : (X, d) → ( X̄ , d̄) and j : (X, d)
→ (X , d ) are both completions of (X, d), then there is a surjective isometry σ :
( X̄ , d̄) → (X , d ) such that σ ◦ i = j.
Proof Define σ on i(X ) by σ (i(x)) = j(x), x ∈ X, Fig. 12.1. This is well-defined as
x = y if i(x) = i(y) and is clearly an isometry. Now we extend σ to an isometry on
X̄ by the previous corollary. The image σ ( X̄ ) is complete, being isometric to the
complete space X̄ , and hence is closed. On the other hand, it contains the dense set
j(X ), so itself is dense. Conclusion: σ ( X̄ ) = X and σ is surjective.
The completion of Q with the standard metric is R. The completion of Q with
the p-adic metric is denoted by Q p . It is locally compact and totally disconnected.
It also carries the structure of a field, called the p-adic number field. It is of central
importance in number theory.
We conclude by looking at an important space of continuous functions.
Definition 12.23 Let X be a locally compact Hausdorff space. A continuous (com-
plex valued) function f on X is said to vanish at infinity if, for any ε > 0, there is a
compact set K in X such that | f (x) |< ε for all x ∈ K c . Let C0 (X ) denote the space
of all continuous functions on X vanishing at infinity and let Cc (X ) be the subspace
of continuous functions with compact support.
Theorem 12.24 Let X be a locally compact Hausdorff space. Then
1) every function in C0 (X ) is bounded;
2) C0 (X ) is complete in the sup metric dsup ;
3) Cc (X ) is dense in C0 (X ).
12.3 Completion 173
12.4.1 Ascoli
12.4.2 Arzelà
Cesare Arzelà (1847–1912) was also an Italian analyst. He studied in Pisa under
Enrico Betti and Ulisse Dini. In 1889 he generalised and clarified Ascoli’s result to
178 12 Complete Metric Spaces
give the Arzelà–Ascoli theorem. Leonid Tonelli was his student. (The formulation
of the theorem for functions on compact metric spaces was given by Fréchet.)
12.4.3 Baire
12.4.4 Banach
Stefan Banach (1892–1945) was the brightest star in the Polish mathematical firma-
ment and was a founding father of the subject of Functional Analysis in the early part
of the twentieth century. Normed spaces which are complete in the metric induced
by the norm are called Banach spaces. These were studied by Banach in his thesis
(completed in 1920, published in 1922), which laid the foundations of functional
analysis. His fixed point theorem that we have discussed also appeared in his thesis.
His 1932 monograph, Théorie des opérations linéaires, was the definitive work on
functional analysis at that time. Banach space, Banach algebra, Hahn–Banach theo-
rem, Banach–Steinhaus theorem, Banach’s closed graph theorem and open mapping
theorem, Banach–Mazur theorem, Banach–Schauder theorem, Banach–Alaoglu the-
orem are all of great importance in functional analysis. Banach–Tarski paradox,
Banach fixed point theorem, Banach–Stone theorem are also well-known. Banach’s
teacher Hugo Steinhaus ‘discovered’ him in a park and took him under his wings.
Steinhaus posed some problems on Fourier series that he was finding difficult and
Banach solved them in a week. Steinhaus used to say that his greatest mathematical
discovery was the discovery of Banach. Several talented youngsters soon gathered
around Banach leading to the golden era of Polish mathematics. Stanislaw Mazur
was a doctoral student of Banach at the University of Lwòw, and there were several
others who were inspired by Banach. During the Nazi occupation of Poland, he was
employed as a lice feeder. He died of cancer in 1945, before the war ended.
References
1. Arzela, C.: Funzioni di Linee. Atti della Reale Academia dei Lincei Rendiconti 5, 342–348
(1889)
2. Bourbaki, N.: General Topology, Chapters 1–4. Springer (1995)
References 179
3. Baire, R.: Sur les fonctions de variables réelles. Ann. di Mat. 3(3), 1–123 (1899). https://doi.
org/10.1007/BF02419243
4. Banach, S.: Sur les opérations dans les ensembles abstraits et les applications aux equations
intégrales. Fund. Math. 3, 133–181 (1922). https://doi.org/10.4064/fm-3-1-133-181
5. Fréchet, M.: Sur quelques point du calcul fonctionnel. Rendiconti di Palermo 22, 1–74 (1906).
https://doi.org/10.1007/BF03018603
6. Ascoli, G.: Le curve limit di una varietà data di curve. Mem. Accad. Lincei 18, 521–586 (1883)
7. Esterle, J.: Mittag-Leffler methods in the theory of Banach algebras and a new approach to
Michael’s problem. In: Greenleaf, F., Gulick, D. (eds.) Proceedings of the Conference on
Banach Algebras and Several Complex Variables, Contemp. Math. Am. Math. Soc. 32 (1984)
pp. 107–129 https://doi.org/10.1090/conm/032
8. Runde, V.: A Taste of Topology. Universitext, Springer (2005)
9. Hausdorff, F. Grundzüge der Mengenlehre. Leipzig (1914)
10. Engelking, R.: General Topology Revised Edition. Heldermann Verlag (1989)
11. Dugundji, J.: Topology. Prentice Hall of India (1975)
12. Kelley, J.L.: General Topology. Affiliated East-West Press (1969)
13. Willard, S.: General Topology. Addison-Wesely (1970)
14. Weierstrass, K.: Über continuirliche Functionen eines reellen Arguments, die für keinen Werth
des letzteren einen bestimmten Differentialquotienten besitzen, Königl. Akad. der Wiss.,
Berlin, Werke 2, pp. 71–74 (1872)
15. Jarnìk, V.: Bolzano and the Foundations of Mathematical Analysis. Society of Czech Mathe-
matics and Physics, Prague (1981)
16. Kowalewski, G.: Über Bolzanos Nichtdifferenzierbare Steige Funktion. Acta Math. 44, 315–
319 (1923). https://doi.org/10.1007/BF02403926
17. Russ, S.: The Mathematical Work of Bernard Bolzano. Oxford University Press (2004)
18. Rychlìk, K.: Theorie der reellen Zahlen im Bolzanos handschriftlichen. Nachlasse, Prague
(1962)
19. Cellérier, C.: Note sur les principes fondamentaux de l’analyse. Bull. des Sci. Math., 2nd series
14, 142–160 (1890)
20. Banach, S.: Über die Baire’sche Katagorie gewisser Functionenmengen. Studia Math. 3, 174–
179 (1931). https://doi.org/10.4064/sm-3-1-174-179
21. Bourbaki, N.: General Topology, Chapters 5–8. Springer (1998)
22. Kuratowski, K.: Topology. Academic Press (1966)
23. Peano, G.: Sur une Courbe qui ramplit Toute une Plane. Math. Ann. 36, 157–160 (1890). https://
doi.org/10.1007/BF01199438
24. Munkres, J.R.: Topology, 2nd edn. Pearson Education, Asia (2001)
25. Sagan, H.: Space-Filling Curves. Springer (1994)
Chapter 13
Combinatorial Methods in Euclidean
Topology
Combinatorial tools and methods have become very powerful in modern mathemat-
ics. Elementary proofs of deep topological results ate obtained here using these.
A central piece is the following higher dimensional intermediate value theorem of
Poincaré.
Soient ξ1 , . . . , ξn fonctions continues de n variables x1 , . . . , xn ; la variable xi est
assujettie à varier entre les limites +ai et −ai . Supposons que, pour xi = ai , ξi soit
constamment positif, et pour xi = ai , constamment négatif; je dis qu’il existera un
système de valeurs des x pour lequel tous les ξ ’s annuleront. (Poincaré 1883) [1, 2]
(“Let ξ1 , . . . , ξn be n continuous functions of n variables x1 , . . . , xn : the variable
xi is allowed to vary between the limits +ai and −ai . Suppose that for xi = ai , ξi is
constantly positive, and that for xi = −ai , ξi is constantly negative; I say there will
exist, a system of values of x for which all the ξ ’s vanish.")
That is the n-dimensional intermediate value theorem of Poincaré. Just as in the
one-dimensional case, this will give the higher dimensional fixed point theorem of
Brouwer.
Some of the most fundamental and hard theorems of topology concern ‘famil-
iar’ geometric objects in the usual Euclidean spaces: curves, balls, spheres and the
Euclidean spaces themselves. Jordan curve theorem in plane topology (that we will
not discuss) is an example of an ‘intuitively obvious’ basic result that is far from
easy to prove. Among such classical results, we consider Poincaré’s higher dimen-
sional intermediate value theorem, Brouwer’s theorems on fixed points and invariance
and Borsuk theorems on maps of the sphere. We present elementary combinatorial
approaches to these and some related results.
Fixed point theorems are important for applications to many areas in mathematics,
game theory, mathematical biology, economics, etc. (e.g. John Nash, Nobel Prize
winner in Economics, 1994, used Kakutani fixed point theorem in his work on game
theory and economics; incidentally, Nash also was a recipient of Abel Prize, 2015, in
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 181
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_13
182 13 Combinatorial Methods in Euclidean Topology
Mathematics for his work on differential topology.) The simplest of these is that any
continuous map f : [0, 1] → [0, 1] has a fixed point (see 1.7), an easy consequence
of the intermediate value theorem. Just as in the one-dimensional case, Poincaré’s
higher dimensional intermediate value theorem is equivalent to a higher dimensional
Brouwer fixed point theorem. The case n = 3 of the fixed point theorem was proved
in 1904 by the Latvian mathematician Bohl [3] (that went unnoticed), and by the
famous Dutch topologist L. E. J. Brouwer in 1909 [4]. The general case was proved
in 1910 by both Hadamard [5] and Brouwer [6]; Brouwer made systematic use
of the recently evolving methods of simplicial complexes introduced by Poincaré.
The standard text book proofs use homology theory or degree theory. The Polish
trio Knaster–Kuratowski–Mazurkiewicz gave a simple proof in 1929 [7], based on
Sperner’s 1928 combinatorial lemma [8]. For some recent developments related to
Sperner’s lemma, see [9, 10] and for those on Brouwer’s theorem, see [11].
In this chapter, we are mainly concerned with cubes and balls. Topologically these
are the same, but in some situations, e.g. combinatorial methods, we need to con-
sider cubes rather than balls. On the other hand, one may prefer balls for geometric
purposes. So we first show that their topological equivalence. More generally, we
show that any compact convex set with interior is homeomorphic to the ball.
A set E in Rn is convex if the line segment joining any two points of E lies
entirely in E. An equivalent condition is that if p1 , . . . , pk are in E, then so is any
convex combination t j p j with t j ≥ 0, t j = 1. The intersection of any collection
of convex sets is convex and so, for any set E, the intersection conv(E) of all convex
sets containing E is the smallest convex set containing E. It is called the convex hull
of E and consists of all convex combinations of points of E.
Lemma 13.1 Let K be a compact convex set in Rn with nonempty interior. Then ∂ K
is homeomorphic to Sn−1 . In fact, if p is an interior point of K , then h : x → x−
x− p
p
is a homeomorphism of ∂ K onto Sn−1 .
Proof By applying the homeomorphism x → x − p, it suffices to consider the case
p = 0; this will give us notational simplification. So assume that 0 ∈ int K and
define h by h(x) = x x
. This is a continuous map Rn \ {0} → Sn−1 . We show that
its restriction to ∂ K is bijective.
To show that h is onto, let y ∈ Sn−1 and τ = sup{t ≥ 0 : t y ∈ K }. This is finite
since K is bounded. τ y ∈ K as K is closed. Since 0 is an interior point of K , some
ball B(0, r ) ⊂ K , and so t y ∈ K for any 0 ≤ t < r , proving τ > 0. By definition,
for every ε > 0, there is a t > τ − ε with t y ∈ / K . This shows that every ball around
τ y contains points not in K . Thus τ y ∈ ∂ K , and clearly h(τ y) = y. This proves that
h : ∂ K → Sn−1 is surjective. See Fig. 13.1.
To prove injectivity, note that the ray joining 0 to y meets the boundary of K at τ y.
Injectivity of h just means that such a point is unique. A ray R from 0 is of the form
13.1 Convex Sets and Balls 183
Corollary 13.3 Let . be any norm on Rn and let Bn and Sn−1 be the corresponding
closed unit ball and the unit sphere, respectively. Then there is a homeomorphism of
Bn onto Bn mapping Sn−1 onto Sn−1 and which preserves antipodal points.
τy = x
t0 y 0
sy0 y=h(x)
ry0 y0 B
K
0 Sn−1
Fig. 13.1 Convex sets with interior are homeomorphic to the ball
184 13 Combinatorial Methods in Euclidean Topology
We are interested in the norms x1 = |x j | and x∞ = max j |x j |. For .1 , the
closed unit ball n = {x ∈ Rn : |x j | ≤ 1} is the convex hull of ±e1 , . . . , ±en . Its
boundary sphere is denoted by ♦n−1 . ♦0 = {±1} and ♦1 is the square with vertices
{(±1, 0), (0, ±1)} in R2 (the shape of our symbol ♦). ♦2 is the octahedron with
equatorial face ♦1 and is made up of eight triangular faces got by joining ±e3 to the
four vertices ±e1 , ±e2 of ♦1 in the (x, y)-plane of R3 . It is got by pasting together the
bases of two pyramids. It will also be convenient to consider ♦n+ = {x ∈ ♦n : xn+1 ≥
0} and ♦n− = {x ∈ ♦n : xn+1 ≤ 0}, the ‘upper’ and ‘lower’ hemispheres. For k < n,
we identify the k-sphere as the subset {x ∈ ♦n : x j = 0, j > k + 1} of the n-sphere.
Note that the (n − 1)-sphere is the ‘equator’ of the n-sphere: ♦n−1 = ♦n+ ∩ ♦n− .
In .∞ , we write n for the ball and n−1 for the sphere. n is the convex hull
of ±e1 ± · · · ± en and is the product [−1, 1]n see Fig. 2.5.
Actually, we will not need the general result 13.2 and we will only need the
previous corollary for .1 and .∞ . These are simple to obtain directly. Recall that
Bn and Sn−1 are the ball and sphere in the Euclidean norm given by x22 = |x j |2 .
A map h on a ball is antipode preserving if h(−x) = −h(x). (±x are said to be
antipodes.)
n = 2, the faces are the four sides of the square. For n = 3, the faces are the six
squares bounding the cube.
We need to consider subdivisions of the cube into smaller subcubes and sub-
cells. For example, the midpoints of the sides of the square and the four vertices
together give a subdivision of the square into 16 triangular cells. We formalise such
subdivisions for any cube.
Consider I n . Let e j be the standard unit n-vectors and write, for k ∈ N, say n >
1, ekj = k1 e j . The combinatorial n-cube is the product C(k) = {0, k1 , . . . , k−1 k
, 1}n .
± k
For brevity, we generally suppress n in the notation for I j , e j , C(k) etc. The faces
of C(k) are C ±j (k) = C(k) ∩ I j± and their union ∂C(k) = ∪ j (C +j (k) ∪ C −j (k), is
the boundary of C(k). Combinatorial subcubes of C(k) are sets of the form Q k =
{c + j a j ekJ : a j ∈ {0, 1}} ⊂ C(k), c ∈ C(k). Analogous definitions and notation
are used for I (a)n : ekj (a) = ak e j , Ca (k) = {0, ±a k
, . . . , ±ka
k
}n , etc.
A combinatorial n-simplex is an ordered set S = v0 , v1 , . . . , vn , with v1 − v0 =
eσk (1) , v2 − v1 = eσk (2) , . . . , vn − vn−1 = eσk (n) , for a permutation σ of {1, . . . , n};
v0 , . . . , vn in C(k) are its vertices. For example, 0, e1k , e1k + e2k , . . . , e1k + · · · + enk
is such an n-simplex with σ = id and one with σ = (1, 2) (the transposition) is
0, e2k , e1k + e2k , . . . , e1k + · · · + enk . The associated geometric n-simplex is the con-
vex hull of the v j :
S = [v0 , v1 , . . . , vn ] := {0n t j v j : t j ≥ 0, t j = 1}.
Note that {v1 − v0 , . . . , vn − v0 } is a basis for Rn and so every x ∈ R is a unique
linear combination x = 0n l j (x)v j with 0n l j = 0. The l j are linear functionals on Rn
and hence are continuous. (If l is a linear functional, l(x) = xi f (ei ) and |l(x)| ≤
lx where l is the norm of (l(e1 ), . . . , l(en )) ∈ Rn . l is uniformly continuous as
|l(x) − l(y)| = |l(x − y)| ≤ lx − y.) C(k) gives rise to a subdivision the cube
into smaller cubes and into n-simplices which become smaller as k gets bigger. The
subset F j (S) = v0 , . . . , v̂ j , . . . , vn , the j-th face of S, is got by omitting the j-th
vertex v j from S. It is not an (n − 1)-simplices as per our definition, unless j = 0
or j = n.
Geometric simplices are points, line segments, (closed) triangular regions, solid
tetrahedra and their higher dimensional analogues. A simplicial complex is a col-
lection of simplices arranged in a natural, nice way. Their study and applications to
geometry and topology was initiated by Poincaré (who else!) in 1899 in his second
paper on Analysis situs. We will not study simplicial complexes, but will only use
cubes and their subsimplices.
Sperner’s Lemma concerns simplices and their subdivisions. To avoid the technical-
ities about simplicial complexes and their subdivisions, we use a form of Sperner’s
Lemma for cubes. The Cubical Sperner Lemma was first given by Kuhn [12] in 1960.
We follow Kulpa’s proof, [13, 14].
v1 S[0] v2
S
S[2] S[1]
v0
Hence, τ
= σ forces τ ( j) = σ ( j + 1) and τ ( j + 1) = σ ( j), so w = w j as in the
definition of S[ j] and S = S[ j]. Uniqueness is thus proved in all the cases.
2) Again, we consider the three cases.
Case i). S = v1 − eσk (1) , v1 , . . . , vn , S[0] = v1 , . . . , vn , vn + eσk (1) and F0 (S) ⊂
C(k). Then the coordinates of all the vertices lie between 0 and 1. To conclude that
j-th coordinate of the vr can not all be 0 or 1 for any j. For j = σ (1) it is clear
that v1 ( j)
= vn ( j) since 0 ≤ v1 ( j) − 1/k, vn ( j) + 1/k ≤ 1. For other coordinates,
since vi+1 − vi = eσk (i) , it follows that vi+1 (σ (i))
= vi (σ (i)). Thus, σ (i) coordinate
of all of v1 , . . . , vn can not be the same and so no coordinate of all the vr can be the
same. In particular, F0 (S) I j± for any j.
Conversely, if F0 (S) is contained in some C ±j , either the j-th coordinates of
v1 , . . . , vn are all 0 or all of them are 1. In either case, from the difference relations
of the vertices, we get that the j-th coordinate of eσk (i) is 0 for i
= 1, so σ (i)
= j
for i
= 1. Hence σ (1) = j. If vi ( j) = 0 for all i, the j-th coordinate of v1 − eσk (1) =
v1 − ekj is −1/k and if vi ( j) = 1 for all i, the j-th coordinate of vn + eσk (1) = vn + ekj
is 1 + 1/k. In any case, not all vertices in S or S[0] have all coordinates in [0,1].
This proves 2) in case i)). The other cases are similar.
1 0 2
F NF
NF NF
1 2 0
F NF
F F
0 0
1
0 2 0 2
0 1 2 1
0 2 1 2
0 1
1 0
Corollary 13.9 Under the assumptions of the cubical Sperner lemma, there is a
fully coloured subcube Ck .
Müger [14] extracted and explicitly stated the following result from Kulpa’s proof
of the Poincaré-Miranda theorem. He called it the higher connectedness of the cube.
The reader may like to ponder over this designation and see what the theorem says
when n = 1.
Theorem 13.11 Let H j± be closed sets in I n such that H j± ⊃ I j± and H j+ ∪ H j− =
I n for each 1 ≤ j ≤ n. Then ∩n1 (H j− ∩ H j+ )
= ∅.
Conclusion: λ(x)
= j − 1 for x ∈ I j+ . Thus, λ|C(k) is a Sperner colouring and, by
the cubical Sperner lemma (13.8), there is a fully coloured n-simplex Sk in C(k) for
each k. Claim: Sk meets each H j± . For, any x ∈ Sk with λ(x) = j ∈ {0, 1, . . . , n}
belongs to F j ⊂ H j+ . If x ∈ Sk and λ(x) = j − 1 ∈ {0, 1, . . . , (n − 1)}, then x ∈/ Fj
/ H j+ or x ∈ I j− . In either case, x ∈ H j− and the claim is proved.
or, equivalently, x ∈
√ previous lemma with X = I , and
n
Now the proof is completed on invoking the
±
{K i } = {Hi }, since diam Sk ≤ diam Ck = n/k → 0.
Proof Write R+ = [0, ∞), R− = (−∞, 0] and apply the theorem (13.11) with
H j± = f j−1 (R± ) to get an x ∈ ∩ j H j± . Then f (x) = 0.
These results hold for any cube [a, b]n in place of I n . For n = 1, the corollary is
precisely the intermediate value theorem.
Corollary 13.14 If g, h : I n → I n are continuous maps and if h(I j± ) ⊂ I j± for all
j, there is a p ∈ I n such that g( p) = h( p).
As mentioned earlier, Brouwer actually stated his theorem for an n-simplex. This is
equivalent to the result for a cube or a ball since there are boundary preserving home-
omorphisms between an n-simplex, an n-cube and an n-ball. The same remarks apply
to Borsuk’s no retract theorem and Bohl’s theorem given below. (See Proposition
13.21.)
For the next corollary, we need a definition. A subset E of a space X is a retract
of X if there is continuous map r : X → E such that r (x) = x for x ∈ E. Such a
13.4 Poincaré and Brouwer 191
map r is called a retraction. This concept, introduced by Borsuk [18], will be studied
in detail in the next chapter. The next corollary is a special case of a result due to
Borsuk [18]; see also 3.6 of [19]. (A main result of [18] is that if B is a retract of Rn ,
the components of Rn \ B are unbounded (Théorème 25, [18]).)
Corollary 13.17 (Borsuk No Retract Theorem, 1931)
The boundary ∂ I n is not a retract of I n .
Exercise We have proved Bohl’s theorem from Borsuk’s theorem and Borsuk’s the-
orem from Brouwer’s theorem. Obtain both the reverse implications. Before we
conclude, we make some simple observations on fixed points and give an application
to matrices.
Definition 13.19 A space X is said to have the fixed point property if every contin-
uous map on X (i.e. X → X ) has a fixed point.
For an extensive treatment of modern fixed point theory, see the thick monograph [20]
of Granas and Dugundji. Shashkin [21] contains a delightful, elementary discussion.
Proposition 13.21 If X and Y are homeomorphic spaces and if X has the fixed point
property, then so does Y .
Example 13.22 The hemisphere {x ∈ Sn : xn+1 ≥ 0} and the positive ‘octant’ Sn+ :=
{x ∈ Sn : x j ≥ 0, 1 ≤ j ≤ n + 1} of the unit sphere are both homeomorphic to the
n-ball Bn and hence have the fixed point property.
Proof A acts linearly on Rn+1 in the canonical way: x → Ax and either of the
assumptions of the corollary ensures that Ax is nonzero and has nonnegative coor-
dinates for x ∈ Sn+ . Thus x → Ax/Ax maps Sn+ to itself and so has a fixed point.
Thus there is an x0 ∈ Sn+ with Ax0 = Ax0 x0 and so Ax0 is a positive eigenvalue
of A.
on invariance of domains and invariance of dimension are among the most basic
theorems in topology and geometry. These were both proved by Brouwer, no doubt
inspired by Poincaré and his methods, in 1911 [6].
One can define dimension for more general spaces. As simple examples, look at
the sphere Sn locally. Points in S1 have neighbourhoods which are arcs and arcs have
dimension one! In fact, an arc is homeomorphic to an interval. Similar arguments
can be given for any Sn and one can say Sn has dimension n. There are theories of
dimension for topological spaces, but we will not discuss them. We will be solely
concerned with Euclidean spaces and show that Rn and Rm are ‘not the same’ as
topological spaces if m
= n. This is the dimension invariance theorem. The domain
invariance theorem is more general.
In this section, first we present a proof due to Müger [14], of a result on dimension
invariance, based on Kulpa’s proof of the cubical Sperner lemma. Then, Kulpa’s
elementary proof of the domain invariance theorem, [22], is given.
A closed set F separates closed sets A and B in X , if there are disjoint open sets
U and V such that A ⊂ U, B ⊂ V and U ∪ V = F c .
We use the earlier notation and results on cubes. Here is a consequence of the
Theorem on Higher Connectedness 13.11.
Proposition 13.24 If a closed set F j in the cube I n separates I j+ and I j− for each
j ∈ {1, . . . , n}, then ∩n1 F j is nonempty.
Proof For each j, there are disjoint open sets U ± ± ± +
j with I j ⊂ U j and F j = U j ∪
c
U− ± ± ± ∓ +
j . The sets H j := U j ∪ F j is closed since X \ H j = U j . For each j, H j ∪
− + − + −
H j = I and 13.11 shows ∩ j (H j ∩ H j ) is nonempty. But H j ∩ H j = F j and so
n
The dimension invariance (13.30) was proved by Brouwer (1911). Lebesgue also
gave proofs of the invariance results. These turned out to be incorrect, but his work
led to the concept of covering dimension.
2+
2−
Fig. 13.5 2
(a) (b)
(c) (d)
Fig. 13.7 a–d Octahedron ♦2 , stages of its barycentric subdivision
second is almost alternating and the last is not even almost alternating. In a symmetric
triangulation aligned to a flag of hemispheres, an alternating or an almost alternating
simplex is admissible if its sign matches that of its support hemisphere.
In the proof of Ky Fan’s Lemma we construct a graph. A (finite) graph G consists
of a finite set V (G) and a subset E(G) ⊂ V (G) × V (G). Points v ∈ V (G) are called
the vertices and elements (v, w) ∈ E(G) are called the edges of G. Geometrically,
an edge is thought of as the line segment joining two vertices (=points). Two vertices
forming an edge are adjacent. They are complementary (for a given labelling ) if
(v) = −(w). The number of edges at a vertex v is called its degree. [v1 , . . . , vr ]
is a path in G if vi , vi+1 are adjacent for 1 ≤ i ≤ r − 1.
13.6 Borsuk and the Sphere 199
Remark 13.36 Look at a triangle and the inscribed circle to conclude n + 1 is the
best possible.
Proof Suppose {U1 , . . . , Un+1 } is an open cover. For x ∈ Sn , there is a j such that
x ∈ U j . Then there is an open ball B(x, r ) with closure B̄(x, r ) ⊂ U j . A finite
number of these open balls cover Sn , by compactness. For each j, let C j be the union
of those closed balls that are contained in U j . Then C j is closed and C j ⊂ U j for
each j and {C1 , . . . , Cn+1 } cover Sn . By the L-S-B theorem above (13.35), some C j
contains a pair of antipodal points and then so does U j .
The results of this section (Borsuk et al.) have several applications in various fields.
See, e.g. [28]. For example, there is a higher dimensional analogue of the dosa
theorem (idli theorem!) usually called the ham sandwich theorem [28, 31, 32].
Exercise 13
hull of the x j , d j := d(x j , κ(x j )), d = κ(x j )d j > 0. Apply Brouwer on K to
f (x) = t j x j , t j = d j /d.)
4. Subdivisions of a 2-simplex. Let S = S0 be an open 2-simplex in R2 and let
Sn+1 be the open 2-simplex with vertices at the midpoints of the sides of Sn , n ≥ 1.
Find ∩Sn .
5. A result of Hadamard, 1910 [5]. Let f : Bn → Rn be continuous and let
f = ( f 1 , . . . , f n ). If j x j f j (x) ≥ 0 for x ∈ Sn−1 , then f (x) = 0 for some x ∈ Bn .
Deduce this from Brouwer’s theorem (13.16). [35]
6. Generalised L-S-B theorem. If Sn is covered by A1 , . . . , An+1 each of which
is either open or closed, then ±x ∈ A j for some x and some j.
7. Real functions on the sphere. If f j , 1 ≤ j ≤ n are real continuous functions
on Sn , there is an x ∈ Sn such that f j (x) = f j (−x) for all j.
8. Integer lattices. Are Zn and Zm homeomorphic for any n, m?
9. Homeomorphisms of balls and spheres. If either Sn and Sm are homeo-
morphic or if Bn and Bm are homeomorphic, then m = n.
10. Dimension invariance from Borsuk–Ulam. Use Borsuk–Ulam (13.34) to
prove that Rn and Rm are not homeomorphic for m
= n. In fact, there is no injective
continuous map Rn → Rm for n > m.
11. Interior and boundary. If E ⊂ Rn and h : E → Rn is a homeomorphic
embedding, then x∈ int E implies h(x)
∈ int h(E) and x ∈ ∂ E implies h(x) ∈ ∂h(E).
12. Embedding a sphere in Euclidean space. (a) Sn has no embedding in Rm
for n ≥ m.
b) Sn is not homeomorphic to any proper subset of itself.
13.7.1 Bohl
13.7.2 Hadamard
13.7.3 Ulam
Stanisław Ulam (1909–1984) Polish mathematician and nuclear scientist got his
doctoral degree in mathematics under Kuratowski. The famous Scottish Café in
Lwów was the place where many famous Polish mathematicians met to discuss
mathematics. Ulam was a member of the Lwów School of Mathematics and an active
contributor to the Scottish Book of problems. In mathematics, he contributed to
several areas like set theory, measure theory, topology, ergodic theory, group theory,
number theory, combinatorics and graph theory. He visited the USA many times on
204 13 Combinatorial Methods in Euclidean Topology
the invitation of von Neumann. During World War II, he worked on the Manhattan
Project at the Los Alamos Laboratory. After the war, he solved the problem of
how to initiate fusion in thermo-nuclear weapons (Teller-Ulam design) and devised
the ’Monte-Carlo method’ widely used in solving mathematical problems using
statistical sampling. Nuclear propulsion, joint work with Enrico Fermi et al. that led
to the creation of nonlinear science, contributions to theoretical biology are some
of the other things he is known for. In mathematics, Borsuk–Ulam theorem, Mazur-
Ulam theorem, Kuratowski-Ulam theorem, Ulam’s packing conjecture are among
his contributions.
13.7.4 Sperner
13.7.5 Mazurkiewicz
13.7.6 Brouwer
dimensions, etc.- and introduced new methods and tools (e.g. simplicial approxi-
mation, degree of a map) that were often compared to a magic wand. His home
near Amsterdam became the focal point where leading topologists (Vietoris, Hopf,
Hurewicz, Alexandroff et al.) met and discussed topology. This burst of brilliance
was perhaps inspired and triggered by Brouwer’s meeting with Poicaré in Paris in
1910 where he also met Borel and Hadamard. For these and more on the remarkable
years, see Freudenthal’s writing in [36] and the article of James [37].
In addition to proving ground-breaking theorems, his methods have also become
standard tools in the subject. He was killed in 1966 in an accident, struck by a vehicle
while crossing the street in front of his house. Brouwer fixed point theorem, Brouwer
degree, Brouwer theorems on invariance of domain and dimension are among those
giving him mathematical immortality.
References
1. Poincaré, H.: Sur certaines solutions particulieres du probleme des trois corps. C. R. Acad.
Sci. Paris 97, 251–252 (1883)
2. Poincaré, H.: Sur certaines solutions particulieres du probleme des trois corps. Bull.
Astronomique 1, 63–74 (1884)
3. Bohl, P.: Über die Bewegung eines mechanischen Systems in der Nähe einer Gleichgewicht-
slage. J. Reine Angew. Math. 127, 179–276 (1904). https://doi.org/10.1515/crll.1904.127.
179
4. Brouwer, L.E.J.: Zur Analysis Situs. Math. Ann. 68, 422–434 (1910) (Collected Works, vol.
2, 354–366.). https://doi.org/10.1007/BF01475781
5. Hadamard, J.H.: Sur quelques applications de l’indice de Kronecker, Introduction to: Tannery.
J. La Theorie des Fonctions d’une Variable, Hermann, Paris (1910)
6. Brouwer, L.E.J.: Beweis der Invarianz der Dimensionenzahl. Math. Ann. 70, 161–165 (1911)
(Collected Works, vol. 2, 430–434.). https://doi.org/10.1007/BF01461154
7. Knaster, B., Kuratowski, K., Mazurkiewicz, S.: Ein Beweis des Fixpunksatzes für n-
dimensionale Simplex. Fund. Math. 14, 132–137 (1929). https://doi.org/10.4064/fm-14-1-
132-137
8. Sperner, E.: Neuer Beweis für die Invarianz der Dimensionszahl und des Gebiets Math. Sem.
Hamburg. Univ. 6, 265–272 (1928). https://doi.org/10.1007/BF02940617
9. De Loera, J.A., Goaoc, X., Meunier, F., Mustafa, N.H.: The discrete yet ubiquitous theorems
of Carathéodory, Helly, Sperner, Tucker and Tverberg Bull. Am. Math. Soc. 56, 415–511
(2019). https://doi.org/10.1090/bull/1653
10. Ivanov, N.V.: Sperener’s lemma, the Brouwer fixed point theorem and cohomolgy (2019).
arXiv: 09065193v2
11. Park, S.: Ninety years of the Brouwer fixed point theorem. Vietnam J. Math. 27, 187–222
(1999)
12. Kuhn, H.W.: Some combinatorial lemmas in topology. IBM J. Res. Devel. 4, 518–524 (1960).
https://doi.org/10.1147/rd.45.0518
13. Kulpa, W.: The Poincaré-Miranda Theorem. Am. Math. Monthly 104, 545–550 (2018). https://
doi.org/10.2307/2975081
14. Müger, M.: A remark on the invariance of dimension. Math. Semesterber. 62, 59–68 (2015)
15. Browder, F.E.: Fixed point theory and nonlinear problems. Bull. Am. Math. Soc. 9, 1–39
(1983). https://doi.org/10.1090/S0273-0979-1983-15153-4
206 13 Combinatorial Methods in Euclidean Topology
16. Poincaré, H.: Sur les courbes définies par une équation difféerentielle IV. J. Math. Pures Appl.
85, 151–217 (1886)
17. Miranda, C.: Un’ osservazione su una teorema di Brouwer. Boll. Unione Mat. Ital. 3, 527
(1940)
18. Borsuk, K.: Sur les rétractes. Fund. Math. 17, 152–170 (1931). https://doi.org/10.4064/fm-
17-1-152-170
19. Borsuk, K.: Theory of Retracts. Monografie Mat, Warsaw (1967)
20. Granas, A., Dugundji, J.: Fixed Point Theory, Monographs in Mathematics. Springer (2003)
21. Shashkin, Y.A.: Fixed Points. American Mathematical Society-Mathematical Association of
America (1991)
22. Kulpa, W.: Poincaré and domain invariance theorem. Acta Univ. Carol. Math. et Phys. 39,
127–136 (1998)
23. Tao, T.: Hilbert’s Fifth Problem and Related Topics. Graduate Studies in Mathematics, vol.
153. Americian Mathematical Society (2014)
24. Borsuk, K.: Drei Sätz über die n-dimensional euklidische Sphäre. Fund. Math. 20, 177–190
(1933). https://doi.org/10.4064/fm-20-1-177-190
25. Tucker, A.W.: Some topological properties of disk and sphere. In: Proceedings of First Cana-
dian Mathematical Congress, Montreal, pp. 285–309. University of Toronto Press, Toronto
(1946)
26. Fan, K.: A generalization of Tucker’s combinatorial lemma with topological applications.
Ann. Math. 2(56), 431–437 (1952). https://doi.org/10.2307/1969651
27. Prescott, T., Su, F.E.: A constructive proof of Ky Fan’s generalization of Tucker’s lemma. J.
Combin. Theory Ser. A 111(2), 257–265 (2005). https://doi.org/10.1016/j.jcta.2004.12.005
28. Matoušek, J.: Using the Borsuk-Ulam Theorem. Springer (2008)
29. Lusternik, L., Schnirelmann, L.: Topological Methods in Variational Problems, (in Russian)
Moscow (1930)
30. Lusternik, L., Schnirelmann, L.: Topological methods in variational problems and their appli-
cations to the differential geometry of surfaces. Uspekhi Mat. Nauk 2(1), 166–217 (1947).
(in Russian)
31. Hatcher, A.: Algebraic Topology. Cambridge University Press (2003)
32. Spanier, E.: Algebraic Topology. Tata-McGraw-Hill (1966)
33. Idzik, A., Kulpa, W., Maćkowiak, P.: Equivalent forms of the Brouwer fixed point theorem I.
Top. Methods Nonlinear Anal. 44, 263–276 (2014)
34. Kuratowski, K.: Topology. Academic Press (1966)
35. Mawhin, J.: Simple Proofs of the Hadamard and Poincaré-Miranda theorems using the
Brouwer fixed point theorem. Am. Math. Monthly 126, 260–263 (2019). https://doi.org/10.
1080/00029890.2019.1551023
36. Brouwer, L.E.J.: In: Freudenthal, H. (ed.) Collected Works, vol. 2, North-Holland (1976)
37. James, I.M.: From combinatorial topology to algebraic topology. In: James, I.M. (ed.) History
of Topology, pp. 561–574. Elsevier (1999)
Chapter 14
Homotopy
etc. (It is interesting to note that study of general topology came into being much later;
e.g. Hausdorff’s classic Grundzüg der Mengenlehre [11] appeared only in 1914; the
definition of a metric appeared first in 1906 in Fréchet’s thesis [12].
We will have a glimpse of algebraic topology (in the form of homotopy) in this
and the next chapter. In the previous chapter we had looked at a bit of combinatorial
topology. For a delightful, lucid invitation to differential topology from the master,
see Milnor [13].
Before looking at homotopy, we discuss the related, useful concepts of retracts and
deformation retracts.
Étant donné un ensemble B ⊂ A, j’appelle fonction rétrancte l’ensemble A en B,
toute fonction f intrinsèque et continue dans A, telle que B = f (A) et qui remplit
l’équation fonctionelle f ( f (x)) = f (x) Borsuk [14].
‘Given a set B ⊂ A, I call a retraction function of A in B every function defined
and continuous on A such that B = f (A) and which satisfies the functional equation
f ( f (x)) = f (x)’.
Thus did Borsuk introduce the concept of retraction in his thesis in 1931. (The
term ‘retract’ was suggested by his thesis advisor Mazurkiewicz.) This and related
concepts have turned out to be quite important in topology. He also obtained most of
the basic properties of retracts that we present here. (See also Borsuk’s monograph
[15].)
Consider the function f defined on R by
⎧
⎨ 0, x ≤ 0
f (x) = x, 0 ≤ x ≤ 1
⎩
1, x ≥ 1
r(x)
x r(x) r(x) x
r(x)
H(t,x)=(1-t)x+tr(x)
r(x)
r(x)
H(t,x)=(1-t)x+tr(x)
‘We shall say that two transformations belong to the same class, if they can be
transformed continuously into each other.’ Brouwer [1]
Homotopy of two maps was thus expressed explicitly for the first time, by
Brouwer. (‘Class’ in Brouwer’s definition is, of course, homotopy class.) Homo-
topy was introduced by Poincaré in his Analysis situs in 1895, but a definition had
to wait till 1912.
Historians of mathematics say that roots of homotopy of maps and paths can
be traced back to Lagrange’s work on calculus of variations and Cauchy’s work
on complex integration, see, e.g., [16]. The term ‘homotopie’ occurs first in Dehn–
Heegard [17]. Brouwer’s definition, in its modern garb:
Definition 14.7 Let X, Y be spaces. If f, g : X → Y are continuous, we say that f
is homotopic to g, written f ∼ g, if there is a continuous map F : X × [0, 1] → Y
such that f (x) = F(x, 0), g(x) = F(x, 1) for every x ∈ X . Such an F is called a
homotopy from f to g and we write F : f ∼ g. F gives a continuous deformation
of f to g.
Another, equivalent, way of looking at homotopy is this: Consider for each
t ∈ [0, 1], the continuous map f t : X → Y defined by f t (x) = F(x, t) for all
212 14 Homotopy
Corollary 14.11 With the notation of the Lemma above, if f ∼ g and h ∼ k, then
h ◦ f ∼ k ◦ g.
Two homeomorphic spaces are ‘topologically equivalent’ or have the ‘same topo-
logical type’. Here is the homotopy counterpart.
Remark 14.13 If X and Y are homeomorphic, then they are homotopically equiva-
lent (but not conversely, as we will see ere long).
Corollary 14.16 A space X is contractible if and only if any two continuous maps
f, g : Y → X are homotopic for every space Y .
Proof Let X be contractible and let x0 , x1 ∈ X . Let c be the constant map c(x) ≡ x1
and let H : id X ∼ c. Then α(s) = H (x0 , s) defines a path in X joining x0 to x1 .
Example 14.18 1. Two null homotopic maps need not be homotopic. In fact, even
two constant maps need not be homotopic. For instance, suppose x0 , x1 belong to two
different path components in X . Then the constant maps c0 (x) = x0 and c1 (x) = x1
are not homotopic.
2. Rn is contractible (Example 14.8). Thus, homotopically speaking, Rn , for n ≥ 1,
is indistinguishable from a point! Thus, homotopy equivalence is a rather crude
relation. It is highly valuable nevertheless.
3. The arguments used for Rn remain valid for any convex set in Rn yielding the
result any convex set in Rn is contractible.
4. Recall the comb space
C̄ = {(1/n, y) ∈ R2 : n ∈ N, 0 ≤ y ≤ 1} ∪ {0} × [0, 1] ∪ [0, 1] × {0}. Define F : C̄
× [0, 1] → C̄ by F((x, y), t) = (x, (1 − t)y). Then F : idC̄ ∼ p1 , the projection on
the x-axis and p1 is homotopic to the constant map 0. So idC̄ is homotopic to 0. Geo-
metrically, push each vertical segment to the x-axis and then push every point on the
x-axis segment to the origin along the x-axis. Thus the comb space is contractible.
214 14 Homotopy
The next couple of results are about homotopy of maps into the sphere and from the
sphere.
Proposition 14.19 Let X be any space. If f, g : X → Sn are continuous maps such
that f (x) and g(x) are not antipodal points for any x ∈ X , then f is homotopic to
g.
(1−t) f (x)+tg(x)
Proof Define F : X × [0, 1] → Sn by F(x, t) = (1−t) f (x)+tg(x)
. Note that the
assumption f (x), g(x) are never antipodal ensures that the denominator never van-
ishes. Then it is clear that F : f ∼ g.
Corollary 14.22 Any continuous map from Sn into a contractible space has a con-
tinuous extension to the closed unit ball Bn+1 .
Corollary 14.24 The identity map Sn → Sn is not null homotopic. Hence Sn is not
contractible.
Proof The identity map is antipodal and so the previous proposition applies. Alter-
natively, if idSn is null homotopic, it has a continuous extension Bn+1 → Sn , contra-
dicting the no retract theorem.
14.2 Homotopy of Maps and Paths 215
a
[0, 1] × {1} b
β
β
x1
[0, 1] × {t} x0
αt
αt
[0, 1] × {0} α
α
Example 14.32 α(s) = (cos π s, sin π s), 0 ≤ s ≤ 1, is a path from (1,0) to (-1,0)
and β(s) = (cos π(s + 1), sin π(s + 1)), joins (-1,0) to (0,1). They are the upper
and lower semicircles, respectively, described in the counter clockwise. In this case,
α
β is a loop at (1,0), i.e. α
β(0) = (1, 0) = α
β(1) and is given by α
β(s) =
(cos 2π s, sin 2π s), 0 ≤ s ≤ 1. It describes a complete circle, starting from (1,0).
x2
id × ϕ.5,1
β βt β
G x1
α αt α
F
ϕ0,.5 × id x0
Similarly H (s, 1) = α
β (s). All these together mean that H : α
β ∼α
˙
β .
The genesis of the proof is that ϕ0,.5 × id : (s, t) → (2s, t) expands the rectangle
[0, 21 ] × [0, 1] to the square [0, 1] × [0, 1] and then F acts on the square; similarly
for [ 21 , 1] × [0, 1] and G (Fig. 14.5).
Now consider three paths in X , α from x0 to x1 , β from x1 to x2 and γ from x2
to x3 . Combining these would yield a path from x0 to x3 . But there are two ways of
combining these: (α
β)
γ and α
(β
γ ). Which is the ‘correct’ way to combine?
Are these related? Let us first write out these explicitly for our edification.
α
β(2s), 0 ≤ s ≤ 21
(α
β)
γ (s) =
γ (2s − 1), 21 ≤ s ≤ 1
⎧
⎨ α(4s), 0 ≤ s ≤ 14
= β(4s − 1), 41 ≤ s ≤ 21
⎩
γ (2s − 1), 21 ≤ s ≤ 1
α(2s), 0 ≤ s ≤ 21
(α
β)
γ (s) =
β
γ (2s − 1), 21 ≤ s ≤ 1
218 14 Homotopy
⎧
⎨ α(2s), 0 ≤ s ≤ 21
= β(4s − 2), 21 ≤ s ≤ 43
⎩
γ (4s − 3), 34 ≤ s ≤ 1
=(α
β)
γ (s)
The last lemma yields the required homotopy Hid,η (Fig. 14.6a).
Definition 14.37 For any point x in a space X , the constant path x at x is defined
by x (s) = x for all s ∈ [0, 1].
14.2 Homotopy of Maps and Paths 219
a b c
α β γ α ex0 α ←
α
α β γ α ex0
Clearly, if we travel along x0 (i.e. we do not move at all, but stand still at x0 ), and
then travel from x0 to x1 along α, effectively we have travelled only along α. Path
homotopy respects this naive description.
Noting that σ (0) = 0, σ (1) = 1, apply 14.34 taking τ = id. In fact, if H (s, t) =
α((1 − t)σ (s) + ts), then H : x0
α ∼α.
˙ Write out the analogous arguments show-
˙ (Fig. 14.6b).
ing α
x1 α
For any path from x0 to x1 , there is a natural reverse path from x1 to x0 ‘described
at the same speed’.
←
Definition 14.39 For a path α from x0 to x1 in a space X , the reverse path α from
←
x1 to x0 is defined by α (s) = α(1 − s), s ∈ [0, 1].
Proof This is easy and predictable: just reverse each member of the family of paths
← ←
deforming α to β - the resulting continuous family of paths deforms α to β . In other
← ← ← ←
˙ then H : α ∼
words, if H : α ∼β, ˙ β , where H (s, t) = H (1 − s, t) for all s, t ∈ [0, 1].
←
Travelling from x0 to x1 along α and then back from x1 to x0 along α is equivalent
to standing still at x0 !
220 14 Homotopy
← ←
Lemma 14.41 If α is a path from x0 to x1 in a space X , then α
α ∼
˙ x0 and α
α ∼
˙ x1 .
←
Proof α
α (s) = α(σ (s)) and x0 (s) = α(τ (s)) where τ ≡ 0 and σ : [0, 1] →
[0, 1] is given by
2s, 0 ≤ s ≤ 21 ,
σ (s) =
2 − 2s, 21 ≤ s ≤ 1,
←
hence H = Hσ τ : α
α ∼˙ x0 (14.34). Explicitly, H (s, t) = α((1 − t)σ (s)). Simi-
larly write down the other part (Fig. 14.6c).
Exercise 14
a metric on P(a, b). Paths α, β ∈ P(a, b) are path homotopic if and only if they lie
in the same path component of P(a, b) (from [18]).
6. Homotopy and fixed points. Is the fixed point property a homotopy invariant?
14.3.1 Borsuk
Karol Borsuk (1905–1982) was a famous topologist from Poland. Retracts were first
defined by him in his thesis in 1931. He also introduced absolute neighbourhood
retracts, which have turned out to be rather important. During World War 2, he
taught at the underground university (along with Kuratowski) in Warsaw. During the
period 1931–1932 he did postdoctoral studies with leading European topologists of
the time, Menger in Wien(=Vienna), Heinz Hopf in Zürich and Vietoris in Innsbruck.
He was the leader of the Warsaw school of topology which was greatly influenced by
his wonderful insight. See [2]. He founded the modern branch of topology known as
shape theory with his celebrated paper of 1968 on homotopic properties of compacta.
He had spent time in leading American universities like Princeton and University of
California, Berkeley. Borsuk–Ulam theorem, Borsuk antipodal theorem and Borsuk
conjecture are well-known in topology. His teacher was the well-known topologist
Mazurkiewicz and the outstanding topologist Samuel Eilenberg was his student.
References
2. James, I.M.: From combinatorial topology to algebraic topology. In: James, I.M. (ed.) History
of Topology, pp. 561–574. Elsevier (1999)
3. Sarkaria, K.S.: The topological work of Henri Poincaré. In: James, I.M. (ed.) History of Topol-
ogy, pp. 123–167. North-Holland (1999)
4. Poincaré, H.: Analysis situs. J. de l’École Polytech. 2(1) 1–123, Oeuvres; vol. VI, 193–288
(1895)
5. Poincaré, H., Complèment à l’Analysis situs, Rend. Circ. Math. d. Palermo, 13, 285–343;
Oeuvres, vol. VI, 290–337 (1899)
6. Poincaré, H.: Second Complèment à l’Analysis situs. Proc. Lond. Math. Soc. 32, 277–308
(1900). https://doi.org/10.1112/plms/s1-32.1.277
7. Poincaré, H., Cinquième Complément à l’Analysis Situs, Rend. Circ. Math. d. Palermo, 18,
45–110; Oeuvres, Vol. VI, 435–498 (1904)
8. Poincaré, H.: Analyse des travaux scientifiques de Henri Poincaré faite par lui-meme (1901).
Acta Math. 38, 1–135 (1921). https://doi.org/10.1007/BF02392063
9. Poincaré, H., Oeuvres, Tome VI, Garniers, R., Leray, J. (eds.) Gauthiers-Villars (1953)
10. Poincaré, H.: Papers on Topology: Analysis Situs and Its Five Supplements. Translated by J.
Am. Math. Soc, Stillwell (2010)
11. Hausdorff, F. Grundzüge der Mengenlehre. Leipzig (1914)
12. Fréchet, M.: Sur quelques point du calcul fonctionnel. Rendiconti di Palermo 22, 1–74 (1906).
https://doi.org/10.1007/BF03018603
13. Milnor, J.: Topology from the Differentiable Viewpoint. University of Virginia Press (1965)
14. Borsuk, K.: Sur les rétractes. Fund. Math. 17, 152–170 (1931). https://doi.org/10.4064/fm-17-
1-152-170
15. Borsuk, K.: Theory of Retracts. Monografie Mat, Warsaw (1967)
16. Dieudonné, J.A.: The work of Nicholas Bourbaki. Am. Math. Monthly 77, 134–145 (1970).
https://doi.org/10.1080/00029890.1970.11992437
17. Dehn, M., Heegaard, P.: Analysis situs. Encyclopedia der mathematischen Wissenschaften mit
Einschluss ihrer Anwendungen, vol. 3, pp. 153–220. Halfte I, Teubner, Leipzig, Teil I (1907)
18. Gamelin, T.W., Greene, R.E.: Topology. Dover (1999)
Chapter 15
Fundamental Groups and Covering
Spaces
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 223
K. Parthasarathy, Topology, La Matematica per il 3+2 134,
https://doi.org/10.1007/978-981-16-9484-4_15
224 15 Fundamental Groups and Covering Spaces
The appearance of Analysis situs marks, by general consensus, the birth of Algebraic
Topology. It contained Poincaré’s construction of homology theory and the funda-
mental group. (For translations of all the parts of Analysis situs, with commentary,
see [3].) The fundamental group, with its remarkable abstract structure, has turned
out to be really of central importance in topology and elsewhere. Here is how the
first definition appeared.
Cela posé, il est clair que l’on peut imaginer un groupe G satisfaisant aux con-
ditions suivantes:
1. A chaque contour fermé M0 B M0 correspondra une substitution S du groupe;
2. La condition nécessaire et suffisante pour que S se réduise à lu substitution
identique, c’est que M0 B M0 ≡ 0;
3. Si S et S correspondent aux contours C et C et si C ≡ C + C , la sub-
stitution correspondant á C sera SS . Le groupe G s’appellera le groupe
fondamental de la variété V .
“La Science dont l’objet est l’étude de ce groupe et de quelques autres analogues
a reçu le nom d’Analysis situs.” [1]
Poincaré’s definition may be translated as follows.
“Given this, it is clear that we can think of a group G satisfying the following
conditions:
1. Each closed contour M0 B M0 corresponds to a substitution S of the group.
2. The necessary and sufficient condition for S to reduce to the identity substitution
is that M0 B M0 ≡ 0.
3. If S and S correspond to contours C and C and if C ≡ C + C then the
substitution corresponding to C is SS . The group G is called the fundamental
group of the manifold V.”
“The science whose object is the study of this group and other analogues get the
name Analysis situs.”
We are going to define the fundamental group essentially as above—as classes of
loops at a point. All the ground work needed for this has been carried out in the last
chapter. We need to just specialise the results on path homotopy obtained there to the
case of closed paths or loops. This immediately leads us to the fundamental group.
Definition 15.1 A path α in a space X is called a loop at x0 if α(0) = α(1) = x0 .
π1 (X, x0 ) is the set of loop homotopy classes at x0 .
Let us examine what we get if the lemmas of the last chapter are applied to loops at
x0 . An alert reader might have already noticed the emergence of a familiar structure
there. Note that if α, β are loops at x0 , then so is α β. For a loop α at x0 , let [α]
denote its loop homotopy class. Thus π1 (X, x0 ) = {[α] : α is a loop at x0 }.
Lemma 14.33 says that if [α] = [α ] and [β] = [β ], then [α β] = [α β ]. This
assures the validity of the following natural definition.
15.1 The Fundamental Group 225
Definition 15.2 Let X be a space and let x0 ∈ X . For [α], [β] in π1 (X, x0 ), define
[α] [β] = [α β].
Let us see what the remaining lemmas say about this binary operation on
π1 (X, x0 ): Lemma 14.36 says that the binary operation is associative. Lemma
14.38 implies that x0 , the constant loop at x0 , acts as the identity element. Finally,
←
the content of Lemma 14.40 is that [ α ], the class of the reverse path of α, is the
inverse of [α] for the operation . The following basic result paraphrases all these
observations.
π1 (X, x0 ) is called the fundamental group of X with base point x0 or of the pointed
space (X, x0 ). The subscript in π1 indicates that perhaps there are more things—the
πn —to come. This is indeed true—one has the homotopy groups πn (X ) and the
fundamental group is just the first of these sequence of groups. However, we will not
be discussing the higher homotopy groups. See [4–6].
One usually considers only path connected spaces while studying fundamental
groups. The following simple result tells us why.
Thus, for a path connected space we can speak of the fundamental group π1 (X )
of the space, without any reference to the base point. Note, however, that the given
isomorphism is not natural and, in general, depends very much on the choice of the
path σ . In fact, it can be shown that the isomorphism σ is independent of σ if and
only if π1 (X, x0 ) is abelian. This last condition puts a severe restriction on X .
Examples of fundamental groups are not easy to come by. Here is the simplest of
these: Any two loops at x0 ∈ Rn are path homotopic and so π1 (Rn , x0 ) is the trivial
one element group. Spaces of this type are important and so have a special name.
226 15 Fundamental Groups and Covering Spaces
Definition 15.5 A simply connected space is a path connected space having a trivial
fundamental group.
The fundamental group gives us a way that leads to groups from spaces: π1 :
(pointed) topological spaces groups. Having a method of passing from one class
of objects (topological spaces) to another (groups), one would like to take along
appropriate maps between objects of the first class to relevant maps between objects
of the second class—in this case, continuous maps between topological spaces to
homomorphisms between their fundamental groups. These are easily got and the
procedure has all the nice properties one would expect.
For convenience, we just write f : (X, x0 ) → (Y, y0 ) to mean that x0 ∈ X,
y0 ∈ Y , f : X → Y and f (x0 ) = y0 .
Proof A contractible space is path connected and has the homotopy type of a point,
whose fundamental group is clearly trivial.
The sphere S2 is simply connected as we will see a little later, but it is a fact that it
is not contractible.
Proof We have proved that a space and its deformation retract are homotopically
equivalent (14.26).
In general, the task of computing the fundamental group of a given space is a difficult
one. As a very simple example, we have seen that Rn has trivial fundamental group.
Determining the fundamental groups of even simple spaces like the circle or the
sphere is nontrivial. With minimal machinery we will try to discuss some simple
examples.
We will revisit this result later in a more general set-up. The case n = 1 is more
involved and we begin with an informal, heuristic discussion about S1 . It is, of course,
path connected. It is intuitively clear that it cannot have trivial fundamental group: the
simple, circular loop (at, say, 1) cannot be shrunk to a point loop. It is also reasonably
‘clear’ that the loop that makes m revolutions is ‘different’ from the one that makes n
revolutions unless m = n. Of course, revolutions can be positive (anti-clockwise) or
negative (clockwise). For example, making a positive revolution and a negative one
will amount to making no revolution at all, i.e., is equivalent to the constant loop.
Thus one can feel that there is a close connection between π1 (S1 ) and integers. We
now proceed to make these intuitive observations precise and rigorous, and identify
the fundamental group of the circle.
The first example of a nontrivial fundamental group that we will see is that of
the circle. The methods involved in this are important and general. However, we
postpone all the technical definitions and concepts involved to the next section. Here
we directly and explicitly work with circle and eschew generalities and jargon.
Proof Take a = b = 0, identify {0} × [0, 1] with [0,1] and take F = α in the basic
lifting lemma. More elaborately, let F : {0} × [0, 1] → S1 be defined by F(0, t) =
α(t) and let F̃ : {0} × [0, 1] → R be its lift. Then α̃(t) = F̃(0, t) gives the required
lifting with α̃(0) = F̃(0, 0) = 1.
Note that even though α is a loop, the lifted path α̃ may not be a loop. The choice
of the base point as 1 is only a convenience. If α is a loop at z 0 = e2πıt0 , it has a
unique lift α̃ as a path in R with initial point t0 and end point t0 + n for some n ∈ Z,
so p ◦ α̃ = α.
The next agendum is to see that homotopies can also be lifted.
Proof We show that the lift F̃ : [0, 1[×[0, 1] → R of F, given by the basic lifting
lemma, is a path homotopy from α̃ to β̃. Observe that s → F̃(s, 0) is a path in R with
initial point F̃(0, 0) = 0 such that p( F̃(s, 0)) = F(s, 0) = α(s). The uniqueness of
15.2 Examples and Applications 231
lifting shows that F̃(s, 0) = α̃(s). Since p( F̃(0, t)) = F(0, t) = 1 for 0 ≤ t ≤ 1,
the continuous function t → F̃(0, t) is integer valued, hence, by connectedness,
F̃(0, t) = F̃(0, 0) = 0, 0 ≤ t ≤ 1. In particular, F̃(0, 1) = 0 and so s → F̃(s, 1) is
the unique lift of β with initial point 0, as argued above for α. That is, F̃(s, 1) = β̃(s).
Moreover, F̃(1, t) is a constant function of t, just as we have proved for F̃(0, t).
Conclusion: F̃ is actually a path homotopy, F̃ : α̃ ∼ ˙ β̃, as asserted.
All these ‘lifting’ results are valid in more generality and these important results will
be taken up in the next section. Now we are ready to enjoy the fruits of our labour in
proving these results: to reach our first goal of identifying the fundamental group of
the circle.
Theorem 15.19 Let α be any loop at 1 in S1 and let α̃ be its unique lift with α̃(0) = 0.
Then
1) deg α := α̃(1) is an integer that depends only on the loop homotopy class of
α, and is called the degree of α.
2) The map deg: π1 (S1 , 1) → Z, deg([α]) := deg α = α̃(1), is an isomorphism
of groups.
The 1-dimensional case of the following theorem was given in Chap. 1, as a conse-
quence of the Intermediate Value Theorem. The general case was given in Chap. 13,
Sect. 6. Here the 2-dimensional case is presented as an application the results of this
section.
Theorem 15.20 (Borsuk-Ulam Theorem, dim 2)
If f : S2 → R2 is continuous, there exists x ∈ S2 such that f (x) = f (−x); i.e. f
takes the same value at some pair of antipodal points.
232 15 Fundamental Groups and Covering Spaces
Proof If the result is false, the function g(x) = f (x) − f (−x) never vanishes and
g(x)
so h(x) = g(x) is a well-defined continuous map S2 → S1 . Define α : [0, 1] → S2
by α(s) = (cos 2π s, sin 2π s, 0); it describes the equatorial circle in S2 . The loop
σ = h ◦ α lifts to a path σ̃ in R. Since h(−x) = −h(x), σ (s + 1/2) = −σ (s) for
0 ≤ s ≤ 1/2, and so σ̃ (s + 1/2) = σ̃ (s) + (2k + 1)/2 for some integer k = k(s).
But k is continuous on [0,1/2] and so must be a constant. Hence
σ̃ (1) = σ̃ (1/2) + (2k + 1)/2
= σ̃ (0) + (2k + 1)/2 + (2k + 1)/2
= σ̃ (0) + (2k + 1).
This shows that σ is not null homotopic. On the other hand, α is null homotopic in
S2 , with null homotopy H say; then h ◦ H is a null homotopy of h ◦ α = σ . This
contradiction completes the proof.
Let us now take a re-look at the sphere. It is intuitively clear that any loop at, say,
x0 = (1, 0, 0) on the sphere S2 can be deformed on the sphere to the single point x0 .
This is also not difficult to prove: Divide such a loop α into small pieces so that each
piece αi lies wholly in either in U = S2 \ {(0, 0, 1)} or V = S2 \ {(0, 0, −1)}. Each
end of αi can be joined to x0 by a path that lies in U or V , as the case may be; these,
together with αi , gives a loop at x0 , lying in U or V . As both U and V are simply
connected (homeomorphic to R2 !), each of these loops is null homotopic. But the
product of all these loops, homotopically, is just α. Conclusion: every loop at x0 is
homotopic to the trivial loop and so S2 is simply connected. These arguments can
15.2 Examples and Applications 233
carried out in a general setting with hardly any extra effort and so we write down the
details in this general context. We begin with the observation that ‘a path is the sum
of its pieces’. We state it as a lemma. The idea of the proof is as in the associative
law for path products.
Proof The case n = 1 being trivial, let us consider the case n = 2. So we have 0 =
t0 < t1 < t2 = 1. Note that α j = α ◦ η j where η1 : [0, 1] → [0, t1 ], η1 (s) = t1 s and
η2 : [0, 1] → [t1 , 1], η2 (s) = (1 − t1 )s + t1 are the respective oriented affine maps.
α1 (2s) = α(η1 (2s)), 0 ≤ s ≤ 21 ,
α1 α2 (s) =
α2 (2s − 1) = α(η2 (2s − 1)), 21 ≤ s ≤ 1.
Corollary 15.23 Take subdivisions 0 = t0 < t1 < · · · < tn = 1 and 0 = s0 < s1 <
· · · < sm = 1 of [0, 1] and let αi and β j be the paths given by α|[ti−1 ,ti ] and α|[s j−1 ,s j ] ,
respectively. Then
(α1 (α2 (· · · ))αn ∼(β
˙ 1 (β2 (· · · ))βm .
collapsed to a single point on the x-y-plane to end up with the punctured plane.
H ((x, y, z), t) := (x, y, (1 − t)z) defines a deformation retraction of X to the x-y-
plane punctured at the origin (see Fig. 14.3) So π1 (X ) is isomorphic to Z.
4. The n-dimensional torus Tn , which is the product of n copies of T := S1 , has
fundamental group Zn , the free abelian group of rank n.
5. The product of two simply connected spaces is simply connected. In particular,
Rn−1 × Sm and Sn × Sm are simply connected for m, n > 1.
6. Rn \ {0} → Sn−1 × (0, ∞), x → (x/x, x), is a homeomorphism. This
shows that Rn \ {0} simply connected for n ≥ 3. For n = 2 we again get that
π1 (R2 \ {0})
π1 (S1 )
Z.
7. Two circles touching externally is a deformation retract of the plane or disc
with two holes (Fig. 14.2). So all these have isomorphic fundamental groups. A circle
together with a diameter (called the theta space) has the same homotopy type, and
hence the same fundamental group, as each of these. But what is this group? We have
not computed it! We only make some heuristic remarks. Think of the picture of two
circles touching externally. Take this point of contact x0 as the base point. You can
form loops at x0 by going around either circle any number of times. These give loops
of the form (with obvious notation) ρ1m 1 ρ2n 1 ρ1m 2 ρ2n 2 · · · ρ1m k ρ2n k , m j , n j
∈ Z. It is true that any loop at x0 is path homotopic to a loop of this form. Thus
the fundamental group consists of elements of the form [ρ1 ]m 1 [ρ2 ]n 1 [ρ1 ]m 2
[ρ2 ]n 2 · · · [ρ1 ]m k [ρ2 ]n k , m j , n j ∈ Z (see [9]). This is the free group on two gen-
erators and is the first nonabelian fundamental group that we have encountered.
Proof The fundamental group of Rn \ {0} is trivial for n > 2 and is nontrivial for
n = 2.
Here are special cases of the general results proved in Chap. 13, but now proved
using fundamental groups.
Proof It suffices to prove the result for monic polynomials. Consider such a polyno-
mial p(z) = z n + a1 z n−1 + · · · + a0 and suppose it has no zeros in C. We can then
define h t (z) := p(t z)/| p(t z)| to get a map h t : S1 → S1 for each t ≥ 0. Note that h 0
is a constant and F(z, s) = h ts (z) gives a continuous map F = Ft : S1 × [0, 1] → S1
for any t > 0. Then F(z, 0) = h 0 and F(z, 1) = h t (z). Thus F is a homotopy from
h 0 to h t . Since h 0 is a constant, it induces the trivial homomorphism on π1 (S1 , 1)
and so h t also induces the trivial homomorphism for all t > 0.
On the other hand, we now show that for sufficiently large t > 0, h t is homotopic to
the map z → z n . We know that this last map induces the nontrivial homomorphism
k → nk and so h t would also induces a nontrivial homomorphism, leading to a
contradiction.
To see the opening assertion of the previous paragraph, observe first that lim h t (z)
t→∞
= z n for each z ∈ S1 . In particular, (1 − s)z n + sh t (z)
= 0, 0 ≤ s ≤ 1, for all large
t. If we let
H (z, s) := (1 − s)z n + sh t (z)/|(1 − s)z n + sh t (z)|,
then H : S1 × [0, 1] → S1 is continuous, H (z, 0) = z n and H (z, 1) = h t (z), thus
giving a homotopy H = Ht from the function z n to h t for each sufficiently large t.
Hence our assertion is proved yielding a contradiction and completing the proof.
their book [11] of 1934, Seifert and Threlfall presented, probably for the first time,
a thorough discussion of the relation between the fundamental group, path lifting
and covering spaces. For more on this, see Dieudonné’s history, [12]. For a more
detailed discussion of fundamental groups and covering spaces, see [5, 9, 10, 13,
14].
Let us begin by looking back at the procedure leading to the identification of
the fundamental group of the circle. It relied on the fact that R and S1 are tied
up topologically in a special way. This relation is implemented by the exponential
map p : R → S1 , p(x) = e2πı x , that was critically used throughout. This surjective
continuous map has some interesting properties. For instance, if U = S1 \ {−1} is the
circle punctured at −1, it is a neighbourhood of 1 in S1 , p−1 (U ) is the disjoint union
∪ (n − 21 , n + 21 ) and p|(n− 21 ,n+ 21 ) is a homeomorphism of this interval onto U for
n∈Z
each n. (For n = 0 this homeomorphism was used in our discussion of π1 (S1 , 1).)
If V is the open semicircular arc in the right half plane joining ı and −ı, then
p−1 (V ) = ∪Vn∗ where Vn∗ = (n − 41 , n + 41 ) and p is a homeomorphism of V onto
Vn∗ . These properties are encapsuled in the concept of covering spaces that is of
fundamental importance in topology and geometry.
X is thought of as the base space and several (finitely or infinitely many) copies of
U stacked above U . What the covering map p does is to just squash down the entire
stack on U . Note that p is a local homeomorphism. (What should the term mean?)
Ũ3
Ũ2
Ũ1 p
X̃
f˜
p
f
Y X
Proof We show that A = {y : f˜(y) = g̃(y)} and B = Ac are both open. Let y ∈ Y
and choose an open neighbourhood U of f (y) that is evenly covered and sheets
Ũ , Ṽ of p −1 (U ) such that f˜(y) ∈ Ũ , g̃(y) ∈ Ṽ . Then Ũ = Ṽ if y ∈ A and they are
disjoint if y ∈ B. By continuity, there is an open neighbourhood V of y such that
V ⊂ f˜−1 (Ũ ) ∩ g̃ −1 (Ṽ ).
Suppose y ∈ A. Then Ũ = Ṽ and for y ∈ V , p( f˜(y )) = f (y ) = p(g̃(y )).
Since p is injective on Ũ = Ṽ , we get f˜(y ) = g̃(y ). So V ⊂ A and hence A is
open. If y ∈ B, then Ũ and Ṽ are disjoint, so f˜(y )
= g̃(y ) for y ∈ V and V ⊂ B,
hence B is open.
15.3 Covering Spaces 239
The underlying idea of the proof of lifting in the general case is the same as in the
special case of R and S1 , but the details are more involved. In that case, we could write
down everything explicitly, but now we have to work with admissible neighbourhoods
and sheets. Cover F([a, b] × [0, 1]) by finitely many admissible open sets Ui and
subdivides [a, b] and [0,1] into small enough intervals I j and Jk such that each
F(I j × Jk ) is contained in a single Ui ; F̃ is then carefully defined piece-wise on
each I j × Jk to give a lift of F| I j ×Jk so that these pieces could be patched up using
the pasting lemma. Once we get the basic lifting, path and homotopy liftings follow.
Proof By uniqueness, we only need to find a lift F̃. If X itself is admissible, take
F̃ = (p|Ũ0 )−1 ◦ F, Ũ0 being the sheet containing x̃0 .
Otherwise, let {Uα } be a cover of X by admissible open sets and let δ > 0
be a Lebesgue number of the open cover {F −1 (Uα )} of Z := [a, b] × [0, 1]. Let
a = s0 < s1 < · · · sm = b, 0 = t0 < t1 < · · · tn = 1 be partitions so that each Ri j =
[si−1 , si ] × [t j−1 , t j ] has diameter less than δ. This ensures that each Ri j is mapped
by F into a single admissible neighbourhood: F(Ri j ) ⊂ Uα for some α. Now F̃ is
defined on Z by first defining it on each Ri j and then pasting all the pieces together.
Begin with the bottom left corner rectangle R11 . F(R11 ) lies in an admissible
neighbourhood U11 of x0 . If Ũ11 is the sheet over U11 with x̃0 ∈ Ũ11 and p(x̃0 ) = x0 ,
then p11 := p|Ũ11 is a homeomorphism onto U11 and F̃11 := (p11 )−1 ◦ F is the lift of
F on R11 with F̃11 (z 0 ) = x̃0 .
Next consider R21 , the rectangle above R11 in the first column. Write z 1 =
(a, t1 ), x̃1 = F̃11 (z 1 ). Choose an admissible neighbourhood U21 of x1 = F(z 1 ) with
F(R21 ) ⊂ U21 and a sheet Ũ21 above it that contains x̃1 . F̃21 := (p21 )−1 ◦ F is the lift
of F on R21 with F̃21 (z 1 ) = x̃1 . Note that x1 ∈ U11 ∩ U21 and if z ∈ F −1 (U11 ∩ U21 ),
then F̃11 (z) = p−1 −1
11 (F(z)) = p21 (F(z)) = F̃21 (z). Thus F̃11 = F̃21 on F (U11 ∩
−1
240 15 Fundamental Groups and Covering Spaces
U21 ) ⊃ R11 ∩ R21 . The pasting lemma yields a continuous map F̃2 with F̃2 = F̃i1
on Ri1 for i = 1, 2. This F̃2 is the lift of F on R11 ∪ R21 with F̃2 (z 0 ) = x̃0 . It is now
clear how to proceed: take the next rectangle R31 in the first column, using admissi-
ble open sets as before, get a lift F̃31 of F on R31 with F̃31 (z 2 = (a, s2 )) = F̃21 (z 2 ).
Paste this with F̃2 to get the lift F̃3 of F on ∪31 Ri1 with F̃3 (z 0 ) = x0 . After m steps
(induction) we get a lift F̃m of F on ∪m 1 Ri1 with F̃m (z 0 ) = x 0 . (See Fig. 15.3.) (U11
and U21 overlap; they both contain the image of the common boundary of R11 and
R21 . For clarity of pictures, we have drawn them separately.)
What we have done so far is this: starting with the bottom left corner rectangle
R11 we have inductively constructed Fi1 on each of the rectangles Ri1 above R11 in
the first column. Similarly, starting with R11 we can move right to get the lifts in each
rectangle R1 j in the bottom row. Again, starting with each bottom rectangle, we can
go up the column and cover all the Ri j . Note that, at each stage Ri j , the lift is chosen
to ‘match with’ the lifts in the adjacent rectangles Ri−1, j and Ri, j−1 , immediately
below and to the left of Ri j . This enables the pasting to be done at each stage. Finally,
after going through the patching at each stage, we get the required lift F̃.
Corollary 15.37 (Path Lifting)
Let p : X̃ → X be a covering map. Let x0 ∈ X and let x̃0 ∈ X̃ be such that p(x̃0 ) =
x0 . If α is a path in X with α(0) = x0 , then there is a unique path α̃ in X̃ with initial
point x̃0 such that p ◦ α̃ = α.
Proof The proof of 15.17 is valid with hardly any changes.
Corollary 15.38 (Homotopy Lifting)
Let p : ( X̃ , x̃0 ) → (X, x0 ) be a covering map. Let α, β be paths in X with α(0) =
x0 = β(0) and α(1) = β(1). Let α̃, β̃ be their lifts with initial point x̃0 . If there is
a path homotopy F : α ∼β, ˙ then there is a unique lift F̃ : [0, 1] × [0, 1] → X̃ of F
with F̃ : α̃ ∼˙ β̃.
Proof Let F̃ be the unique lift of F with F̃(0, 0) = x̃0 . Then
p( F̃(s, 0)) = F(s, 0) = α(s), p( F̃(s, 1)) = F(s, 1) = β(s).
By uniqueness of liftings, F̃(s, 0) = α̃(s), F̃(s, 1) = β̃(s). Further,
p( F̃(0, t)) = F(0, t) = x0 and p( F̃(1, t)) = F(1, t) = α(1) = β(1).
Thus F̃(0, t) is in the discrete space p−1 {x0 }, so F̃(0, t) ≡ F̃(0, 0) = x̃0 , by conti-
nuity and F̃ : α̃ ∼
˙ β̃. In the same way, F̃(1, t) is a constant.
Proposition 15.39 Let p : X̃ → X be a covering map and let x0 ∈ X , x̃0 ∈ X̃ with
p(x̃0 ) = x0 . Then p : π1 ( X̃ , x̃0 ) → π1 (X, x0 ) is injective; in particular, if X is sim-
ply connected, then so is X̃ .
Proof Suppose α̃ is a loop at x̃0 in X̃ and F : p ◦ α̃ ∼e
˙ x0 . Let F̃ be the lift of F such
that F̃(0, 0) = x0 . Then F̃ : α̃ ∼e
˙ x̃0 .
Proposition 15.40 Let p : ( X̃ , x̃0 ) → (X, x0 ) be a covering map. If X̃ is path con-
nected, then [α] → α̃(1) is a well-defined map η : π1 (X, x0 ) → X̃ , where α̃ is the
lift of α with α̃(0) = x̃0 . Further
15.3 Covering Spaces 241
Corollary 15.41 The fundamental group of the real projective space is the cyclic
group of order 2: π1 (RPn )
Z2 , n > 1.
Proof p : Sn → RPn , x → [x] = {±x}, is a covering. Each p−1 (x0 ) has two ele-
ments and Sn is simply connected; 2) applies.
Proof [α] ∈ H ⇔ [α] ∈ H [ex0 ] ⇔ α̃(1) = ẽx0 (1) = ex̃0 (1) = x̃0 .
∼
Proof Let x̃0 ∈ p−1 (x0 ), x1 ∈ p−1 (x1 ) and let H j = p (π1 ( X̃ , x̃ j )). It suffices, by
2) of 15.42, to give a bijection between the right coset spaces π1 (X, x0 )\H0 and
π1 (X, x1 )\H1 . If σ̃ is a path from x̃0 to x̃1 in X̃ , then σ := p ◦ σ̃ is a path in X from
←
x0 to x1 . If σ ([α]) = [ σ α σ ] and if [α] = [p ◦ α̃] ∈ H0 , then
∼ ∼ ∼
← ← ←
σ ([α]) = [p ◦ σ p ◦ α̃ p ◦ σ̃ ] = [p ◦ ( σ α̃ σ̃ )] = p ([ σ α̃ σ̃ ]) ∈ H1 .
←
Thus σ maps H0 to H1 and its inverse σ maps H1 to H0 , hence σ (H0 ) = H1 and
H0 [α] ↔ H1 σ ([α]) is the required bijection.
Corollary 15.45 If x̃0 , x̃1 ∈ p−1 (x0 ), then H0 := p (π1 ( X̃ , x̃0 )) and H1 :
= p (π1 ( X̃ , x̃1 )) are conjugate subgroups in π1 (X, x0 ).
Proposition 15.48 Suppose X has a simply connected covering ( X̃ , p). Then every
x ∈ X has a neighbourhood U such that each loop at x in U is path homotopic in
X to the constant loop ex .
We conclude the chapter and the book with the basic result on the existence of
simply connected cover. For proofs, see [5, 9, 14].
Theorem 15.50 Let X be a space which is path connected, locally path connected
and semilocally simply connected and let x0 ∈ X .
1) For any subgroup H of π1 (X, x0 ), there is a path connected covering p :
( X̃ , x̃0 ) → (X, x0 ) such that p (π1 ( X̃ , x̃0 )) = H . This covering is unique: If q :
(Ỹ , ỹ0 ) → (X, x0 ) is a covering such that q (π1 (Ỹ , ỹ0 )) = H , then there is an iso-
morphism h : ( X̃ , x̃0 ) → (Ỹ , ỹ0 ).
2) In particular, X has a unique simply connected covering space.
The simply connected covering space X̃ guaranteed by the theorem is called the
universal covering space of X . For example R is the universal cover for S1 , Sn for
RPn , n ≥ 2, and Rn for Tn , n ≥ 1.
Exercise 15
15.4.1 Poincaré
variables. I had need of it for the study of periods of multiple integrals and for the
application of this study to the development of the perturbation function. Finally I
glimpsed in Analysis Situs a means of attacking an important problem in the theory
of groups, the search for discrete or finite groups contained in a given continuous
group. It is for all these reasons that I devoted to this science a fairly long work.”)
His collected works in topology is [17]; English translation may be found in [3]. His
writings had to keep pace with the constant pouring of ideas in various areas. He
relied more on his great intuition than rigorous proofs. See [18].
Poincaré conjecture, Poincaré duality, Poincaré-Brirkhoff-Witt theorem, Poincaré-
Hopf theorem, Poincaré metric, Poincaré upper half-pane, Poincaré group, Poincaré-
Bendixon theorem, Poincaré recurrence theorem, Poincaré-Hilbert series are some
of the mathematical objects named after him. In France, a mathematical institute, a
mathematical journal, a university are all named after him. A moon crater and an
asteroid are also given his name.
It would be nice to conclude with a mention of the Poincaré conjecture. In Analysis
situs [19] he conjectured that if a 3-manifold M has the same homology groups as the
sphere S3 , then it is homeomorphic to S3 . But later, in his fifth complement [2], he gave
a counter example—a homology 3-sphere with a nontrivial fundamental group) and
gave the revised (correct) conjecture: Every simply connected closed 3-manifold
is homeomorphic to S3 . After almost a century, this was solved in the affirmative by
the Russian mathematician Grigori Perelman in 2002–2003 (using analytic methods
of Ricci flow developed by Richard Hamilton). The higher dimensional conjectures
had been solved earlier: Steve Smale for n ≥ 5 in 1961 and for n = 4 by Michael
Freedman in 1982. Smale, Freedman and Perelman were chosen for Fields medal—
Perelman declined it.
Here is how Poincare stated it: (V is the 3-manifold under consideration.) “One
question remains to be dealt with: Is it possible for the fundamental group of V
to reduce to the identity without V being simply connected” (i.e. without being
homeomorphic to S3 ?) He concludes with the words: “However, this question would
carry us too far away.”
We sign off with the original in French: “Il resterait une question à traiter: Est-il
possible que le groupe fondamental de V se réduise à la substitution identique, et que
pourtant V ne soit pas simplement connexe? ... Mais cette question nous entraînerait
troploin.”
References
1. Poincaré,H., Sur l’analysis situs, C. R. Acad. Sci. 118, 663–666, Oeuvres; vol. VI, 189–192
(1892)
2. Poincaré, H., Cinquième Complément à l’Analysis Situs, Rend. Circ. Math. d. Palermo, 18,
45–110; Oeuvres, Vol. VI, 435–498 (1904)
3. Poincaré, H.: Papers on Topology: Analysis Situs and Its Five Supplements. Translated by J.
Am. Math. Soc, Stillwell (2010)
4. Gamelin, T.W., Greene, R.E.: Topology. Dover (1999)
References 247
To the reader: Do not look at these until you have tried hard enough. Sufficient
investment of time and thought on a problem is important. Even if you fail to get a
solution, the effort you have put in will yield its own fruits. Suggestions given here
may not be the only, nor even the best, line of attack. You can possibly come up with
a different (and perhaps better) way of doing things. Doing is learning! So, get back
to thinking!
Chapter 1
2(d) Apply IVT to x 2 − a on a suitable interval.
2(f) Suffices to show: if b lies between f (0) and f (1), say f (0) < b < f (1),
then there is an a between 0 and 1 such that f (a) = b. g(t) = bt − f (t) has a
maximum on [0,1] and the maximum cannot be attained at 0 or 1. So there is 0 <
a < 1 with g (a) = 0.
3(a) Draw graphs.
3(b) Think geometrically.
Chapter 2
Chapter 5
1(a) Take R with the topology T = {V : 0 ∈ V } ∪ {∅}. Then {0} is compact, but
what is its closure?
Take Z with the usual (discrete) topology and X = Z ∪ {±1/2}. Consider the
topology on X so that neighbourhoods of 1/2 are X and Z ∪ {1/2}. Similarly for −1/2.
Then K + = Z ∪ {1/2} and K − = Z ∪ {−1/2} are compact, but their intersection is
not.
1(c) Use f.i.p. for the collection {K i ∩ V c }.
1(d) Show that C is closed and bounded. For boundedness, note that uniform
convergence implies f (K ) ∪ (∪∞ n f (K m )) is bounded for large n. To show that C
is closed, take yk ∈ C, yk → y. If infinitely many yk lie in a single f n (K ), we are
done. If not, going to subsequences wherever needed, yk = f n k (xk ) with xk → x.
Then y = f (x).
3(a) δ = d(x0 , f (X )) > 0, ∃xn , d(xn , xm ) ≥ δ, m = n, xn+1 = f (xn ).
4(b) Map a line to a generator or wind it around as a helix.
7(a) Find a space in which no two nonempty open sets are disjoint.
7(c) Here is an elementary proof. If X is not compact, there is a sequence {xn } of
distinct elements with no limit points. Suppose the xn are separated: inf{d(xn , xm ) :
n = m} > δ > 0. For any x, there is at most one n with d(x, xn ) < δ/2. Define
f (x) = n(1 − 2d(x, xn )/δ) if there is such an n. Otherwise define f (x) = 0. Then
f is continuous and f (xn ) = n.
Suppose inf{d(xn , xm ) : n = m} = 0. Then there is a subsequence {xn k } with
limk,l d(xn k , xnl ) = 0, hence g(x) = limk d(x, xn k ) exists and nonzero for all x. But
g(xnl ) → 0. Thus f (x) = 1/g(x) is continuous and unbounded.
The more common proof defines f (xn ) = n for all n and invokes Tietze’s exten-
sion theorem (Chap. 12) to extend f to X .
7(d) d(x, y) + | f (x) − f (y)| with f as in (c).
11. Try finding a convergent subsequence of the sequence {en } of standard unit
vectors.
14. Heine–Borel–Lebesgue again. K is bounded implies the set σ (K) is bounded.
That it is closed follows from two facts: (1) every sequence in K has a convergent
subsequence; (2) the set of singular matrices is closed.
252 Appendix: Selected Exercises—Suggestions and Hints
Chapter 6
Chapter 8
2(c) If E = {xn } has no limit points, then E is closed. For each n there is a
neighbourhood Vn of xn , with xk ∈ / V, k = n. The Vn and E c form a countable open
cover without a finite subcover.
If there is a countable open cover {Vn } without a finite subcover, choose xn ∈ / ∪n1 Vk .
Can the set of xn have a limit point?
3(c) X be infinite, Hausdorff and E the set of isolated points. If E is finite,
inductively get {xn } in E c and open Vn with xn ∈ Vn and xk ∈ / Vn for k < n. xn ∈ /
{xk : k = n} and {xn } is infinite discrete.
5(a) If V is open, x ∈ V , to show: there is a continuous f with x ∈ {y : f (y) =
0} ⊂ V . This is almost immediate from the definition.
5(b) Assume f ≥ 0. To show: for each x0 ∈ X with f (x0 ) > 0 and each 0 < a <
f (x0 ), there is a continuous g ≤ f such that g(x0 ) ≥ a. There is a neighbourhood U
of x0 such that f (x) ≥ a on U . Choose a continuous function g such that g(x0 ) = a,
g = 0 on U c and 0 ≤ g ≤ a.
7(c) Write A = ∩Vn , Vn open, V n+1 ⊂ Vn with B ⊂ V1 . Apply Urysohn for An , B
c
for each n to get an f n . Take f = f n /2 . For the second part, get a g for B as in
n
Chapter 9
2(a) Is {(x, sin(1/x)) : 0 < x < 1} ∪ {(0, 0)} connected?
2(e) Take a ∈ Ac , b ∈ B c and (ξ, η) ∈ X × Y \ (A × B). Find a connected set
containing (a, b) and (ξ, η). Use the following facts: (i) X × y and x × Y are con-
nected for all x ∈ X, y ∈ Y . (ii) The union of connected sets with nonempty inter-
section is connected.
3. A punctured disc is connected.
4(c) (0,1] is a component of B, not homeomorphic to any component of A.
f (1) = 2, f (x) = x otherwise. Map (0, 1] ∪ (3, 4) in B to (0,1) in A and translate
all the other Bn to Bn − 3 to get g.
4(d) If a, b ∈ A, a < b, take a < c < b. A = ((−∞, c) ∩ A) ∪ ([c, ∞) ∩ A) is
a disconnection.
A continuous image of a connected space is connected.
6(d) Let C be a connected subset of a countable metric space X . Take a ∈ C. The
function f (x) = d(x, a) is a continuous function on C and so f (C) is a countable
connected set of real numbers.
Chapter 10
1(a) What is the subspace topology on the x-axis?
1(b) Separable. First countable, even second countable if X is countable and
not first countable if X is uncountable. Compact, hence Lindelöf. What about the
co-countable topology?
2(e) Recall Weierstrass approximation theorem.
2(f) For ∞ , observe that the set of sequences of zeros and ones is uncountable
and the distance between any two of these is 1. For the last part, what can you say
about the set of characteristic functions of subsets of X ?
2(h) The set D of all functions with finite range is dense. Find a countable subset
of D which is also dense.
2(i) Spaces in which no two nonempty open sets are disjoint?
3(d) Can you find an uncountable discrete subset? (It is also compact. Can you
prove this using the result that in any totally ordered set with lub property, any closed
interval with the order topology is compact (which is proved in [1]).)
4(d) The closed set {1/n : n ∈ N} and 0 cannot be separated.
4(e) As X is second countable, we need only to prove: X is sequentially compact
if and only if every countable open cover has a finite subcover. Suppose {Vn } is an
open cover with no finite subcover. Take xn ∈ X \ ∪n1 Vk . {xn } has no convergent
subsequence. For the converse, if a sequence {xn } has infinitely many distinct points,
it has a limit point and there is a subsequence converging to this limit point.
6(f) X regular, Lindelöf, A, B disjoint, closed. {U : U open, Ū ∩ B = ∅} is a
cover for A and a similar open cover for B, These and X \ (A ∪ B) form an open
cover for X . So there is a countable subcollection Un , Ūn ∩ B = ∅, covering A and
an analogous collection {Vn } covering B. Make them disjoint as in the proof of 10.13
and take the unions.
Appendix: Selected Exercises—Suggestions and Hints 255
Chapter 11
1. Is [a, b) compact in R ? Is the unit ball in 2 compact?
2(c) What is the one-point compactification of [0, 1) ∪ (1, 2]?
4. Use the function f (x) = e−1/(a−x) for x > a and 0 otherwise to get a smooth
function on R that vanishes outside [a, b]. Use this, in turn, to get a smooth function
that is 1 for x < a and 0 for x > b (a < b). This yields a smooth function on Rn that
is 1 on a ball and vanishes outside a bigger ball. Use such functions.
5. Use smooth Urysohn to get smooth partitions of unity, just as we did in the con-
tinuous case. (Reference: suitable advanced calculus books or differential topology
books.)
6. For x ∈ f −1 (K ), let Vx be a precompact neighbourhood. f (Vx ) form an open
cover of K , so finitely many V j = Vx j cover K . A = ∪V̄ j is compact and K ⊂ f (A).
C = A ∩ f −1 (K ) is compact and f (C) = K .
7(b) What are the neighbourhoods of the point [Z] in the quotient?
8(a) Let f (x) = lim f n (x), f n continuous. For each m, n, let Vmn be the interior
of the set ∩k≥m {x : | f m (x) − f k (x)| ≤ 1/n}, Wn = ∪m≥1 Vmn and let E = ∩n≥1 Wn .
f is continuous at each point of E, E c is of first category and the points of continuity
of f is dense. (E is open, dense.)
8(b) f is continuous at x if and only if f −1 (In (x)) is an open set containing x
for all n, where In (x) = ( f (x) − 1/n, f (x) + 1/n). Thus the set C( f ) of points of
continuity of f is ∩Vn , where Vn is the union of all open intervals In with length
(In ) < 1/2n.
8(c) No countable dense set in R is a G δ .
Chapter 12
2(b) C[−1, 1] is d1 -dense in L [−1, 1].
1
5(d) Suppose X is complete and let {An } be a Cauchy sequence of closed bounded
sets in the Hausdorff metric d H . Let A be the set of limits of all sequences {an } in
X with an ∈ An for all n. Show that A is the d H -limit of {An }. For this, it is to be
shown that, for any ε > 0, An ⊂ A(2ε) and A ⊂ An (2ε) for large n.
Take ε > 0, fix N with d H (An , Am ) < ε m, n ≥ N . A ⊂ A N (2ε).
Next, choose Nk < Nk+1 for all k such that d H (Am , An ) < ε/2k for m, n ≥ Nk .
Observe that for each a ∈ An and k with n ≥ Nk , there is a bm ∈ Am such that
d(a, bm ) < ε/2k if m ≥ n ≥ Nk .
Using the observation, for each a ∈ An with n ≥ N1 , inductively construct a
sequence {xl } with xl ∈ Al such that d(xi , x j ) < ε/2k for Nk ≤ i, j ≤ Nk+1 and
d(a, x N2 ) < ε/2. Then {x j } is a Cauchy sequence in X and converges to a b with
d(a, b) < 2ε. An ⊂ A(2ε) if n ≥ N1 .
6(b) If (X, d) is not compact, it is not countably compact and there is a decreasing
sequence {Fn } of closed sets with
empty intersection. Define f n (x) = d(x, Fn ) for
each n and define ρ(x, y) = 2−n ρn (x, y) where ρn (x, y) = [| f n (x) − f n (y)| +
256 Appendix: Selected Exercises—Suggestions and Hints
(min{ f n (x), f n (y)})d(x, y)]. ρ is a metric that is equivalent to d and diam ρ (Fn ) ≤
2−n . If xn ∈ Fn , {xn } is a ρ-Cauchy sequence and has no limit as ∩Fn is empty.
6(d) Let ε > 0 and let {x1 , . . . , xn } be an ε-mesh for X . Let {A j }21 −1 be the
n
Chapter 13
5. Let r be the retraction of Rn onto the unit ball and take g = r − f ◦ r . Then g
maps a suitable ball into itself. Apply Brouwer to g. (See Mawhin [2].)
9. Stereographic projection.
10. Assume n > m and h : Rn → Rm is continuous and injective. Consider the
restriction h|Sn−1 .
12(a) Borsuk–Ulam.
12(b) Sn \ { p} is homeomorphic to Rn .
Chapter 14
1(b) Let X be Haudorff and r : X → A be a retraction. If x ∈ Ā, x ∈ / A, there are
disjoint neighbourhoods U of x and V of r (x) with r (U ) ⊂ V . There is a y ∈ A ∩ U
and a = r (a) ⊂ V , a contradiction.
2(a) When are two constant maps homotopic?
2(b) If f has an extension F to Cone(X ), use the quotient map q and F to get a
null homotopy. For the converse, use the homotopy and the fact that q restricted to
X × {0} is a homeomorphism.
4. What does the given condition say about the line segment [α(s), β(s)]?
Chapter 15
2(f) Move each point of the cone to the vertex along a generator.
4(b) Move f (x) to g(x) via the vertex, along generators.
5. Use 4(c).
Remark on vector fields on spheres. A vector field v on Sn is a tangent vector
field if v(x) ⊥ x for all x ∈ Sn . It is very easy to see that there are nonvanishing
tangent vector fields on S2n−1 for all n. A famous result due to Poincaré (1885) says
this is not the case on S2 . This was later extended to all even dimensional spheres.
This is known as the hairy ball theorem. It says that any tangent vector field on
Appendix: Selected Exercises—Suggestions and Hints 257
S2n necessarily vanishes at some point. (‘You can’t comb a porcupine’!) Milnor’s
beautiful proof of this is in [3]. For S2 , a simple proof using Sperner’s lemma is given
in [4].
7(c) iv) Is it simply connected?
7(c) vi) Plane with a hole.
9(b) What about outward radial projection?
11. What is the relation between the annulus and S1 × (a, b)?
References
D
Darboux’s theorem, 5 F
Deck transformation, 242 Face, 185
Dedekind, Richard, 6, 56, 75 Facet, 198
Deformation retract, 210 Fermat, Pierre, 56
Index 263
G
Galois, Evariste, 54, 140 I
Gauß/Gauss, Karl Friedrich, 22, 37, 56 Identification topology, 85
Generalised sequence, 43 Idli theorem/ham sandwich theorem, 201
General topology, 21 Inadequacy of sequences, 31, 43
264 Index
K
Kelley, John, 44, 70, 83, 99, 104 M
Killing, Wilhelm, 56 Matheron, Étienne, 96
KKM map, 202 Matrix space, 18
Klein bottle, 90, 93 Maximal element, 100
Klein, Felix, 56, 90, 93 Mazurkiewicz, Stefan, 141, 182, 204, 208,
Knaster, Bronisław, 135, 141, 182, 204 221
Knaster-Kuratowski-Mazurkiewicz theo- Mazur, Stanislaw, 178
rem/KKM theorem, 202 Menger, Karl, 140
Kolmogorov, Andrei, 160 Method of bisection, 60
Kovalevskaya, Sofia, 55 Metric, 8
Kronecker, Leopold, 56, 74 on a product, 9, 89, 173
Kulpa, Władysław, 193 Metric from a norm, 18
Kuratowski closure operation, 36 Metric space, 8
Kuratowski, Kazimierz/Casimir, 36, 96, 135, compactness, 64
140, 141, 176, 182, 204, 221 Metric topology, 23
and Zorn’s lemma, 101 Metrisable space, 23, 146
characterisation of compactness, 95 Milnor, John, 208, 257
Ky Fan Lemma, 198 Minimal element, 100
Minkowski, Hermann, 56
Mittag-Leffler, Gösta, 56, 167
L Mittag-Leffler’s Theorem, 168
Labelling, 187, 198 Möbius, August Ferdinand, 90, 92
Lattice, 24 Möbius band, 90, 92
Lebesgue covering lemma, 64 Montel, Paul, 74, 161
Lebesgue, Henri, 61, 64, 67, 74, 196, 204 Moore, Robert Lee, 92
Lebesgue number, 64 Moore-Smith convergence, 44
Lebesgue number lemma, 64 Motion group of Rn , 51
Lennes, N.J., 127, 131 Müger, Michael, 193
definition of connectedness, 127 Multiplicative linear functional, 102
Index 265
N Pinching a subspace, 88
Nash, John, 182 Plücker, Julius, 93
Neighbourhood, 11, 25 Poincaré conjecture, 223, 246
Neighbourhood axioms, 22 Poincaré, Henri, 21, 208, 245
Neighbourhood base, 26 and fixed point theorems, 189
Net, 43 and Relativity, 245
Nets vs filters, 53 definition of fundamental group, 224
Nonabelian fundamental group, 235 Gaston Darboux on, 245
Non-archimedean metric, 8 Hairy Ball Theorem, 257
Noncompact balls, 73 higher dimensional intermediate value
Noncontractibility of the sphere, 214 theorem, 181
No Retract Theorem, 191, 201, 202, 214, 235 inspiration to Brouwer, 192
Norm, 10, 19 on geometry and fundamental group, 223
Normal space, 111 on his own work, 245
Normed algebra, 102 on Weierstrass, 56
Nowhere dense set, 153, 169 Poincaré-Miranda theorem, 190
Null homotopic, 213 Pointed topological space, 226
Points of continuity, 160
Polish space, 176
O Pontryagin, Lev, 160
Octahedron ♦n , 184 Positive eigenvalue, 192
Odd degree polynomials, 3 Positive octant of the sphere, 192
One-point compactification, 155 Precompact/Relatively compact, 152
Open ball, 11 Product topology, 28, 83
Open base, 26 Čech’s definition, 79
Open map, 83 Tychonoff’s defintion, 79
Open sets, 22, 23 Projection, 83
Orbit, 89 Projective limit, 168
Orbit space, 89 Projective line, 90
Projective space, 90
Order topology, 34
Projective system, 168
Oriented affine map, 216
Proper map, 96, 101, 159
Orthogonal group O(n), 89
Pseudocompact space, 72
Orthogonal matrix, 89
Pugh, Charles, 98
P Q
Paracompact space, 122 Quotient map, 84, 86
Partial cluster point, 98 Quotient topology, 84
Partial map, 97
Partial net, 98
Partial order, 97, 100 R
Partial product, 97 Raman-Sundström, Manya, 61
Partition of unity, 123, 158 Regular closed set, 36
Pasting lemma, 47 Regular open set, 36
Path, 128 Regular space, 116
Path connected component, 129 Relative topology, 23
Path connected space, 128 Retract, 86, 191, 208
Path homotopy, 215 Retraction, 86, 208
Path Lifting Theorem, 230, 240 Reverse path, 128, 219
Peano curve, 123 Riemann, Bernhard, 21, 37, 56
Perfectly normal space, 122 and topology, 21
Pfaff, Johann, 92 creator of topology, 21
Picard, Émile, 203 Riesz, Frigyes, 21, 38, 68, 107, 127, 131
266 Index
U
Ulam, Stanisław, 203 W
Ultrafilter, 73 Weak* topology, 102
Ultranet/Universal net, 73 Weaker topology, 24
Uniform continuity, 65, 66
Weierstraß/Weierstrass, Karl, 6, 7, 41, 42, 55
and Dirichlet, 65
Poincaré on, 56
and Heine, 65
Weil, André, 36, 102
Uniform continuity theorem, 67
Well-ordered set, 100
Uniformity, 36
Base for a, 37 Well-ordering, 100
Topology given by a, 36 Well-ordering Principle, 100
Uniformly equivalent metrics, 174 Weyl, Hermann, 21, 37
Uniform space, 36
Uniqueness of lifts, 238
Universal cover, 243
Upper bound, 100 Z
Upper half-plane, 91 Zariski topology, 33
Upper hemi-sphere, 184 Zermelo-Fräenkel set theory, 100, 101
Urysohn metrisation theorem, 146 Zorn’s lemma
Urysohn, Pavel, 68, 103, 112, 113, 117, 124, and Bochner, 101
146, 147, 160 and Kuratowski, 101
Urysohn’s Lemma, 113 Zorn, Max, 101, 141
for locally compact spaces, 158 Zorn’s lemma, 97–101, 141
Usual topology, 23 Zygmund, Antoni, 204