You are on page 1of 337

Visual Geometry

and Topology
Anatolij Fomenko
as guest Professor
at the Mathematical Institute
in Heidelberg
Anatolij Fomenko

Visual Geometry
and Topology
With 50 Full-page Illustrations
and 287 Drawings

Springer-Verlag
Berlin Heidelberg New York
London Paris Tokyo
HongKong Barcelona
Budapest
Author:
Anatolij T.Fomenko
Dept. of Differential Geometry and Applications
Faculty of Mathematics and Mechanics
Moscow University, Moscow 119899
Russia

Translator:
Marianna V. Tsaplina
Moscow, Russia
Title of the Russian edition:
Naglyadnaya geometriya i topologia
Moscow University Press, 1993 (abridged version)

Mathematics Subject Classification (1991):


49-XX, 51-XX, 53-xx, 55-XX, 57-XX, 58-XX, 70-XX, 81-XX, 83-XX

Library of Congress Cataloging-in-Publication Data


Fomenko, A. T. Visual geometry and topology/ Anatolij Fomenko;
[translator, Marianna V. Tsaplinaj. p. em.
Includes bibliographical references and index.
ISBN-13: 978-3-642-76237-6 e-ISBN-13: 978-3-642-76235-2
DOl: 10.1007/978-3-642-76235-2
1. Geometry. 2. Topology. I. Title. QA445.F58 1993 516-dc20 92-39676 CIP
This work is subject to copyright. All rights are reserved, whether the whole or part of
the material is concerned, specifically the rights of translation, reprint~, reuse of
illustrations, recitation, broadcasting, reproduction on microfilms or in any other way,
and storage in data banks. Duplication of this publication or parts thereofis permitted
only under the provisions of the German Copyright Law of September 9,1965, in its
current version, and permission for use must always be obtained from Springer-
Verlag. Violations are liable for prosecution under the German Copyright Law.
© Springer-Verlag Berlin Heidelberg 1994
Softcover reprint of the hardcover 1st edition 1994
Typesetting: Springer TEX in-house system

41/3140 - 5 4 3 2 1 0 - Printed on acid-free paper


Preface

Modem geometry and topology take a special place in mathematics be-


cause many of the objects they deal with are treated using visual methods.
At the same time, these visual methods now successfully undergo for-
malization and far-reaching abstraction which have contributed much to
the remarkable progress in modem geometry and its applications. David
Hilbert wrote in 1932: "As far as geometry is concerned, the tendency
of abstraction in it had led to grand systematic constructions of alge-
braic geometry, Riemann geometry and topology where the methods of
abstract reasoning, symbolism and analysis are widely used. But non the
less, visual perception still plays the leading role in geometry, and not
only for being strongly demonstrative in the course of investigation, but
also for the understanding and estimation of the results obtained in the
course of investigation" [l].
Many geometric concepts arose from concrete problems of mecha-
nics, physics, etc., and we shall point out here some of these links. For
important modem mechanisms of the appearance of topological ideas in
the framework of classical mechanics and mathematical physics see, for
example, the papers by Smale [2], Poston and Stewart [3], Hildebrandt
and Tromba [4], Novikov [5], Arnold [7], Nitsche [8]. The mathematical
life is now being actively intruded by computer geometry which permits,
in particular, visualization of intricate mechanical objects that result from
long computational experiments and whose geometric character is hardly
perdictable. It is relevant to mention here the remarkable works by Ban-
choff [9], Mandelbrot [10], Francis [11], Penrose [12], Poston & Stewart
[3], and Peitgen & Richter [13]. Some geometrical aspects of the the-
ory of probability and mathematical statistics are elucidated in the book
by Kolmogorov [14], B.V. Gnedenko [15], Shiryaev [16] and Chentsov
[17]. Some of these ideas are reflected in the sections "Visual Material"
of the present book.
VI Preface

We have selected only the fragments of geometrical knowledge


which are most visual and closely related to applications. See, in par-
ticular, Chap.3 devoted to some modem visual aspects of symplectic
topology and Hamiltonian mechanics. See also the contributions due to
Maslov [18], Faddeev & Zakharov [19], Gelfand [20], Matveev [46], [47],
Zieschang [65], Bolsinov [70] and Kozlov [21].
The principal aim of the present book is to narrate, in an accessible
and fairly visual language, about some classical and modem achieve-
ments of geometry in both intrinsic mathematical problems and applica-
tions. We do not restrict our consideration to the classics, and also touch
upon current problems which are being rapidly developed today. We lay
special emphasis upon visual explanation of the statement of problems,
the methods of their solution and the results obtained and try to acquaint
the reader with geometric ideas as soon as possible, ignoring for the
time being the abstract logical aspect of calculations, considerations, etc.
After having read this book, the reader will be able himself to compre-
hend, when reading specialized literature, more formal approaches to the
problems pointed out in this book. Many modem fields of mathematics
admit visual presentations which do not, of course, claim to be logically
rigorous but, on the other hand, offer a prompt introduction into the sub-
ject matter. In this connection, in Chap. 1 we give a brief presentation
of the classical theory of polyhedra and simplicial homologies because
these ideas are now widely used in mathematics, physics, etc., but their
logical simplification and perfection is often achieved by a higher level
of abstraction. In this respect, the geometric approach to the theory of
homologies, which goes back to Poincare, is perhaps more cumbersome
(in what concerns theoretical grounds) but appreciably simpler and more
visual (and therefore more comprehensive at early stages of acquaintance
with the subject).
Geometrical intuition plays an essential role in contemporary algebro-
topological and geometric studies. Many profound scientific mathemat-
ical papers devoted to multi-dimensional geometry use intensively the
"visual slang" such as, say, "cut the surface", "glue together the strips",
"glue the cylinder", "evert the sphere", etc., typical of the studies of two-
and three-dimensional images. Such a terminology is not a caprice of ma-
thematicians, but rather a "practical necessity" since its employment and
the mathematical thinking in these terms appear to be quite necessary for
Preface VII

the proof of technically very sophisticated results. It happens rather fre-


quently that the proof of one or another mathematical fact can at ftrst be
"seen", and only after that (and following this visual idea) can we present
a logically consistent formulation, which is sometimes a very difficult
task requiring serious intellectual efforts. The expediency of these efforts
is however psychologically justifted by the visual and beautiful picture
already created in the head of the researcher and convincing him that he
had taken the right way. Thus, the criterion of beauty of one or another
geometric image often serves as a compass for choosing an optimal way
of a further formal logical proof. For a more thorough acquaintance with
these ideas the reader may be referred to the well-known books by Manin
[22], [23], where many interesting details can be found.
It was not our aim to give a systematic presentation of individual
ftelds of geometry, but we started on a journey' round its rich world
and on our way "took photos" of those fragments which we thought of
as particularly interesting. The visual and scientiftcally urgent material
being exceedingly abundant, a complete or a systematic presentation is
out of the question. What we offer is a short "diary", an attempt to
narrate to a wide range of mathematicians, mechanics, physicists about
the diversity of methods and applications of modem geometry, to help
them recognize really exciting geometric and physical objects against
the background of sophisticated abstractions. Each chapter of the book
is written as autonomously as possible, so that the reader could plunge
into the ideas and concepts of each section and choose for himself the
order of reading separete chapters.
The author is greatful to S.V. Matveev, Ya.V. Tatarinov, A.V. Cher-
navsky, E.B. Vinberg, V.O. Bugaenko and A.A. Zenkin who kindly pre-
sented some materials on visual geometry and topology.
Each mathematician has his own system of concepts of the intrin-
sic geometry of his (speciftc) mathematical world and visual images
which he associated with some or other abstract concepts of mathematics
(including algebra, number theory, analysis, etc.). It is noteworthy that
sometimes one and the same abstraction brings about the same visual
picture in different mathematicians, but these pictures born by imagina-
tion are in most cases very difficult to represent graphically, so to say,
to draw. Part of the graphical material contained in the present book
is an attempt to ''photograph from within" the sophisticated, peculiar
VIII Preface

mathematical world generously endowed with images and concepts that


constitute the subject matter of contemporary. These graphical represen-
tations are compiled in sections called "Visual Material". Almost each
figure of such a section is endowed with a comment in the text. Some
of the figures complement the content of the chapters with concepts not
reflected in the main text. In this case, such figures give references to
the corresponding literature. Our graphical material is either based on
concrete geometric constructions, ideas, theorems and depict real math-
ematical objects and processes or reflects various ways of perception of
mathematical concepts, for instance, infinity, homotopy, etc. The first at-
tempt in this direction was the book "Homotopic Topology" by Fomenko,
Fuchs and Gutenmacher [24]. This topic was further developed in a new
book "The Course in Homotopic Topology" by Fomenko & Fuchs [25].
The reader can also see the author's art book, Mathematical Impres-
sions [26], which contains high-quality reproductions of approximately
80 of the author's works, including some in colour. The album also
contains short mathematical and extra-mathematical comments to the
pictures. Because they were written with different purposes, these two
books complement each other.
Different sections of the present book are intended for different lev-
els of mathematical knowledge, but the greater part of the material is in-
tended for the first- and second-year students of mathematics or physics.
The book includes some visual aspects of the results in the field of
modem computer geometry obtained in the framework of the scientific
research "Computer Geometry" headed by AT. Fomenko at the Faculty
of Mechanics and Mathematics of Moscow State University. These re-
sults have been discussed at the scientific seminar "Computer Geometry"
working at the Department of Differential Geometry and Applications,
Department of Higher Geometry and Topology and Department of Com-
putational Methods headed by AV. Bolsinov, V.L. Golo, I.Kh. Sabitov,
E.G. Sklyarenko, V.V. Trofimov and AT. Fomenko in Moscow State
University.
The readers who is interested in the details of modem geometri-
cal methods can continue his education using the books by Dubrovin,
Fomenko, Novikov [27] and the book by Fomenko [28].
Preface IX

The book is intended for natural sciences students (beginning from


the first year), post-graduates and specialists interested in applications of
modem geometry and topology.
The book included 50 graphical sheets drawn by the author and pre-
sented in sections Visual Material. The majority of them are made in
pencil and Indian ink on paper and exhibit half tones and sophisticated
light and shade technique. The preparation of precise and high-quality
photocopies, which the author submitted to Springer-Verlag for reproduc-
tion in the book, was therefore a rather difficult task. It was successfully
fulfilled by a professional photographer N.S. Moiseenko (Moscow) to
whom the author expresses his gratitude.
The author is deeply indebted to Springer Publishers.
Table of Contents

1 Polyhedra. Simplicial Complexes. Homologies

1.1 Polyhedra . . . . . . . . . . . . . . . . . . . . . . . . .
1.1.1 Introductory Remarks . . . . . . . . . . . . . . . . . . . .
1.1.2 The Concept of an n-Dimensional Simplex
Barycentric Coordinates . . . . . . . . . . . . . . . . . . 5
1.1.3 Polyhedra. Simplicial Subdivisions of Polyhedra.
Simplicial Complexes . . . . . . . . . . . . . . . . . . .. 8
1.1.4 Examples of Polyhedra . . . . . . . . . . . . . . . . . . . 10
1.1.5 Barycentric Subdivision . . . . . . . . . . . . . . . . . . 14
1.1.6 Visual Material ., . . . . . . . . . . . . . . . . . . . . . 16

1.2 Simplicial Homology Groups of Simplicial Complexes


(polyhedra) . . . . . . . . . . . . . . . . . . . . . . . .. 18
1.2.1 Simplicial Chains . . . . . . . . . . . . . . . . . . . . .. 18
1.2.2 Chain Boundary . . . . . . . . . . . . . . . . . . . . . . 23
1.2.3 The Simplest Properties of the Boundary
Operator Cycles. Boundaries ..... . . . . . . . . . . . 26
1.2.4 Examples of Calculations of the Boundary Operator . . . . 27
1.2.5 Simplicial Homology Groups . . . . . . . . . . . . . . . . 29
1.2.6 Examples of Calculations of Homology Groups.
Homologies of Two-dimensional Surfaces . . . . . . . . . 32
1.2.7 Visual Material .,. . . . . . . . . . . . . . . . . . . . . 46

1.3 General Properties of Simplicial Homology Groups .. 49


1.3.1 Incidence Matrices . . . . . . . . . . . . . . . . . . . . . 49
1.3.2 The Method of Calculation of Homology Groups
Using Incidence Matrices . . . . . . . . . . . . . . . . . . 50
XlI Table of Contents

1.3.3 "Traces" of Cell Homologies Inside Simplicial Ones ... 56


1.3.4 Chain Homotopy. Independence of Simplicial Homologies
of a Polyhedron of the Choice of Triangulation ...... 59
1.3.5 Visual Material . . . . . . . . . . . . . . . . . . . . . . . 65

2 Low-Dimensional Manifolds

2.1 Basic Concepts of Differential Geometry . . . . . . . . 75


2.1.1 Coordinates in a Region.
Transformations of Curvilinear Coordinates . . . . . . . . 75
2.1.2 The Concept of a Manifold. Smooth Manifolds.
Submanifolds and Ways of Defining Them. Manifolds
with Boundary. Tangent Space and Tangent Bundle . . . . 79
2.1.3 Orientability and Non-Orientability. The Differential
of a Mapping. Regular Values and Regular Points.
Embeddings and Immersions of Manifolds.
Critical Points of Smooth Functions on Manifolds.
Index of Nondegenerate Critical Points
and Morse Functions . . . . . . . . . . . . . . . . . . . . 85
2.1.4 Vector and Covector Fields. Integral Trajectories.
Vector Field Commutators. The Lie Algebra
of Vector Fields on a Manifold . . . . . . . . . . . . . . . 91
2.1.5 Visual Material . . . . . . . . . . . . . . . . . . . . . . . 95

2.2 Visual Properties of One-Dimensional Manifolds . . . . 98


2.2.1 Isotopies, Frames . . . . . . . . . . . . . . . . . . . . . . 98
2.2.2 Visual Material ...... . . . . . . . . . . . . . . . . . 102

2.3 Visual Properties of Two-Dimensional Manifolds . . . . 111


2.3.1 Two-Dimensional Manifolds with Boundary . . . . . . . . 111
2.3.2 Examples of Two-Dimensional Manifolds . . . . . . . . . 113
2.3.3 Modelling of a Projective Plane
in a Three-Dimensional Space . . . . . . . . . . . . . . . 115
2.3.4 Two Series of Two-Dimensional Closed Manifolds .... 120
Table of Contents XIII

2.3.5 Classification of Closed 2-Manifolds . . . . . . . . . . . . 125


2.3.6 Inversion of a Two-Dimensional Sphere . . . . . . . . . . 128
2.3.7 Visual Material ., . . . . . . . . . . . . . . . . . . . . . 130

2.4 Cohomology Groups and Differential Forms ...... 135


2.4.1 Differential I-Forms on a Smooth Manifold . . . . . . . . 135
2.4.2 Closed and Exact Forms on a Two-Dimensional Manifold 136
2.4.3 An Important Property of Cohomology Groups ...... 138
2.4.4 Direct Calculation of One-Dimensional Cohomology
Groups of One-Dimensional Manifolds . . . . . . . . . . 139
2.4.5 Direct Calculation of One-Dimensional Cohomology
Groups of a Plane, a Two-Dimensional Sphere
and a Torus . . . . . . . . . . . . . . . . . . . . . . . . . 141
2.4.6 Direct Calculation of One-Dimensional Cohomology
Groups of Oriented Surfaces, i.e. Spheres with Handles . . 148
2.4.7 An Algorithm for Recognition of Two-Dimensional
Manifolds. Elements of Two-Dimensional
Computer Geometry . . . . . . . . . . . . . . . . . . . . 152
2.4.8 Calculation of One-Dimensional Cohomologies
of a Surface Using Triangulation . . . . . . . . . . . . . . 154
2.4.9 Visual Material . . . . . . . . . . . . . . . . . . . . . . . 154

2.5 Visual Properties of Three-Dimensional Manifolds .. . 159


2.5.1 Heegaard Splittings (or Diagrams) . . . . . . . . . . . . . 159
2.5.2 Examples of Three-Dimensional Manifolds . . . . . . . . 162
2.5.3 Equivalence of Heegaard Splittings . . . . . . . . . . . . 164
2.5.4 Spines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
2.5.5 Special Spines . . . . . . . . . . . . . . . . . . . . . . . 168
2.5.6 Filtration of 3-Manifolds with Respect to Matveev's
Complexity . . . . . . . . . . . . . . . . . . . . . . . . . 170
2.5.7 Simplification of Special Spines . . . . . . . . . . . . . . 173
2.5.8 The Use of Computers in Three-Dimensional Topology.
Enumeration of Manifolds in Increasing Order
of Complexity . . . . . . . . . . . . . . . . . . . . . . . 177
2.5.9 Matveev's Complexity of 3-Manifolds
and Simplex Glueings . . . . . . . . . . . . . . . . . . . 184
2.5.10 Visual Material . . . . . . . . . . . . . . . . . . . . . . . 188
XIV Table of Contents

3 Visual Symplectic Topology


and Visual Hamiltonian Mechanics

3.1 Some Concepts of Hamiltonian Geometry . . . . . . . . 193


3.1.1 Hamiltonian Systems on Symplectic Manifolds ...... 193
3.1.2 Invo1utive Integrals and Liouville Tori . . . . . . . . . . . 196
3.1.3 Momentum Mapping of an Integrable System ..... . . 200
3.1.4 Surgery on Liouville Tori at Critical Energy Values .... 200
3.1.5 Visual Material . . . . . . . . . . . . . . . . . . . . . . . 204

3.2 Qualitative Questions of Geometric Integration


of Some Differential Equations. Classification
of Typical Surgeries of Liouville Tori
of Integrable Systems with Bott Integrals . . . . . . . . 208
3.2.1 Nondegenerate (Bott) Integrals . . . . . . . . . . . . . . . 208
3.2.2 Classification of Surgeries of Bott Position
on Liouville Tori . . . . . . . . . . . . . . . . . . . . . . 210
3.2.3 The Topological Structure of Critical Energy Levels
at a Fixed Second Integral . . . . . . . . . . . . . . . . . 215
3.2.4 Examples from Mechanics. The Equations of Motion
of a Rigid Body. The Poisson Sphere.
Geometrical Interpretation of Mechanical Systems . . . . . 216
3.2.5 An Example of an Investigation of a Mechanical System.
The Liouville System on the Plane . . . . . . . . . . . . . 219
3.2.6 The Liouville System on the Sphere . . . . . . . . . . . . 220
3.2.7 Inertial Motion of a Gyrostat . . . . . . . . . . . . . . . . 221
3.2.8 The Case of Chaplygin-Sretensky . . . . . . . . . . . . . 223
3.2.9 The Case of Kovalevskaya ...... . . . . . . . . . . . 224
3.2.10 Visual Material . . . . . . . . . . . . . . . . . . . . . . . 225

3.3 Three-Dimensional Manifolds and Visual Geometry


of Isoenergy Surfaces of Integrable Systems . . . . . . . 231
3.3.1 A One-Dimensional Graph as a Hamiltonian Diagram ... 231
3.3.2 What Familiar Manifolds Are Encountered
Among Isoenergy Surfaces? . . . . . . . . . . . . . . . . 234
Table of Contents XV

3.3.3 The Simplest Isoenergy Surfaces (with Boundary) ..... 243


3.3.4 Any Isoenergy Surface of an Integrable Nondegenerate
System Falls into the Sum of Five (or Two) Types
of Elementary Bricks . . . . . . . . . . . . . . . . . . . . 244
3.3.5 New Topological Properties of the Isoenergy
Surfaces Class . . . . . . . . . . . . . . . . . . . . . . . 246
3.3.6 One Example of a Computer Use in Symplectic Topology 249
3.3.7 Visual Material . . . . . . . . . . . . . . . . . . . . . . . 252

4 Visual Images in Some Other Fields of Geometry


and Its Applications

4.1 Visual Geometry of Soap Films. Minimal Surfaces . . . 255


4.1.1 Boundaries Between Physical Media. Minimal Surfaces .. 255
4.1.2 Some Examples of Minimal Surfaces . . . . . . . . . . . 258
4.1.3 Visual Material . . . . . . . . . . . . . . . . . . . . . . . 260

4.2 Fractal Geometry and Homeomorphisms . . . . . . . . 264


4.2.1 Various Concepts of Dimension . . . . . . . . . . . . . . 264
4.2.2 Fractals . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
4.2.3 Homeomorphisms...................... 268
4.2.4 Visual Material .,. . . . . . . . . . . . . . . . . . . . . 276

4.3 Visual Computer Geometry in the Number Theory .. 284

Appendix 1 Visual Geometry of Some Natural


and Nonholonomic Systems
1.1 On Projection of Liouville Tori in Systems
with Separation of Variables . . . . . . . . . . . . . . . . 293
1.2 What Are Nonholonomic Constraints? ..... . . . . . . 295
1.3 The Variety of Manifolds in the Suslov Problem ..... 297
XVI Table of Contents

Appendix 2 Visual Hyperbolic Geometry


2.1 Discrete Groups and Their Fundamental Region . . . . . . 300
2.2 Discrete Groups Generated by Reflections in the Plane .. 301
2.3 The Gram Matrix and the Coxeter Scheme . . . . . . . . 303
2.4 Reflection-Generated Discrete Groups in Space ...... 303
2.5 A Model of the Lobachevskian Plane . . . . . . . . . . . 306
2.6 Convex Polygons on the Lobachevskian Plane . . . . . . . 308
2.7 Coxeter Polygons on the Lobachevskian Plane . . . . . . . 309
2.8 Coxeter Polyhedra in the Lobachevskian Space ...... 310
2.9 Discrete Groups of Motions of Lobachevskian Space
and Groups of Integer-Valued Automorphisms
of Hyperbolic Quadratic Forms . . . . . . . . . . . . . . . 314
2.10 Reflection-Generated Discrete Groups
in High-Dimensional Lobachevskian Spaces . . . . . . . . 315

References ............................ 317


1. Polyhedra. Simplicial Complexes. Homologies

1.1 Polyhedra

1.1.1 Introductory Remarks

Geometry and topology most often deal with geometrical figures, objects
realized as a set of points in a Euclidean space (maybe of many dimen-
sions). It is useful to view these objects not as rigid (solid) bodies, but
as figures that admit continuous deformation preserving some qualita-

'~
tive properties of the object. Recall that the mapping of one object onto I
I ~
another is called continuous if it can be determined by means of continu- I __

ous functions in a Cartesian coordinate system in space. The mapping of ---


//-- ~~~~-~-

one figure onto another is called homeomorphism if it is continuous and


one-to-one, i.e. establishes a one-to-one correspondence between points
of both figures. In this case there exists the inverse mapping with the
same properties. For example, a straight line segment and a continuous
arc (without self-intersections) on a plane are homeomorphic (Fig. 1.1.1).
Homeomorphic are also a square and a circle (Fig. 1.1.1), a cube and a
tetrahedron (sometimes called a simplex) (Fig. 1.1.1), a plane and a sphere
with one punctured (discarded) point (Fig. 1.1.1). In the latter case the Fig.1.1.1
homeomorphism can be realized using the so-called stereographic pro-
jection (Fig. 1.1.2). To this end one should put a standard sphere onto
a plane, take the point of tangency for the south pole and the very top
point for the north pole N. Then one draws a ray from the north pole
through an arbitrary point x on the sphere and continue the ray up to
its intersection with the horizontal plane. The point obtained will be de-
noted by f(x). The map x - t f(x) just determines the homeomorphism
between the sphere without a point and the Euclidean plane. Fig.1.1.2
2 1. Polyhedra. Simplicial Complexes. Homologies

The concept of homeomorphism appears to be convenient for es-


tablishing those important properties of figures which remain unchanged
under such deformations. These properties are sometimes referred to as
Topological and topological, as distinguished from metrical, which are customarily asso-
metric properties ciated with distances between points, angles between lines, edges of a
figure, etc. A cube and a tetrahedron (Fig.1.l.1) are of course different
from the metric point of view, but at the same time they are homeo-
morphic. For many problems, minute metrical properties of a figure are
Embeddings often inessential, and therefore it is of interest to display its more "rough"
topological properties which are usually masked.
Sometimes, homeomorphism of two figures is not immediately ob-
vious. The simplest example is an ordinary flat circle 8' in a three-
dimensional Euclidean space R3 and a knotted circle (Fig. 1.1.3). They
are of course homeomorphic. To establish this fact, it suffices to cut
each of these curves at one point and superpose the two segments ob-
tained. The image of these figures in space seems different only due to
the fact that the circle is embedded in R3 differently (in two ways). In a
three-dimensional space one cannot continuously and disjointly deform
a nontrivial knot into a flat circle. In a four-dimensional space one can
already do so. A usual strip (a ring) and a strip twice wound in a three-
dimensional space (Fig.l.1.3) are homeomorphic in exactly the same
manner. We invite the reader to prove it himself.
One should not think, however, that any winding of a strip (with any
number of loops) leads to a homeomorphic figure. We shall examine, for
Fig. 1.1.3 example, a strip wound only once (Fig. 1.1.3). It can be verified that such
a strip is already not homeomorphic to a flat ring (to a usual strip). A
Mobius strip strip wound only once is called the Mobius strip (or the Mobius band).
or Mobius band The question is: What will happen to a strip if we wind it any even or
odd number of times?
A set of points homeomorphic to a straight line segment we shall
Simple arc call a simple arc. A simple arc contains two distinguished points, the
ends. They are sometimes called arc vertices. Under a homeomorphism
of a simple arc onto itself, its vertices are carried into themselves or
into one another. The rest of the points of the simple arc (the so-called
interior points) are equivalent in the sense that for any pair x and y of
interior points of the arc there always exists a homeomorphism of the arc
onto itself, such that the point x is mapped into the point y. The visual
1.1 Polyhedra 3

picture of this is given in Fig. 1.1.4 as a graph. Clearly, in one of the


two cases depicted the arc ends exchange places, while in the other case
they go into themselves, i.e. remain in their own places.
A zero-dimensional simplicial complex (a zero-dimensional polyhe-
dron) is a finite set of points (the vertices of the polyhedron). A one-
dimensional complex or graph (a one-dimensional graph) is a set consist-
ing of a finite number ao of points (''vertices'') and a finite number aj
of simple arcs ("edges"). Given this, the following two properties should
be satisfied.
1) Any two edges either have no common points or have only one
common end-point. Each vertex either does not belong to a single edge ~ :f
(such vertices will be called "isolated") or serves as the end-point of one Fig. 1.1.4
or several edges. The number of such edges is called the degree or the
index of the vertex.

o
2) Both end-points of each edge are included in the number of
vertices of the graph (Fig. 1.1.5).
A graph is called connected if it cannot be divided into two subgraphs
without common vertices and edges. Otherwise the graph is disconnected
and falls into several connected subgraphs called components of this
graph (connected components). r_r'

b-l
We have begun with the concept of a graph in order that already at
the beginning of Chap. 1 we could visually demonstrate the basic con-
cepts which are further on developed in a more general case of arbitrary
polyhedra. The reader who will now "feel" these concepts will easily
orientate in the entire material of the book. r -f'
If we remove one edge from the graph r without removing a single Fig. U.S
vertex (including the end-points of the edge), we shall have a subgraph
r'. This operation may apparently vary the number of connected com-
ponents in the graph. The number of components of the graph r' will
remain the same (as in the graph r) if the removed edge joined the ver-
tices of one and the same component of r ' (Fig. 1.1.5). On the contrary, Components of the graph
the number of components of the graph r will increase by one if the
removed edge joined vertices of different components of the graph r'.
The order of connectedness of a graph is the maximal number of
edges which can be removed without changing the number of components
of the graph.
4 1. Polyhedra. Simplicial Complexes. Homologies

A simple cycle is a connected graph all of whose vertices have index


2. It is depicted as a circle. More precisely, such a graph is homeomorphic
to a circle (Fig. 1.l.6).
It is not true that each graph can be realized as a set of points
on a plane. For example, the graph in Fig. 1.1.6 (the set of edges of a
tetrahedron) cannot be embedded into a plane. By the way, this graph is
not a simple cycle (verify this!).
We shall soon get acquainted with the general concept of homolo-
Fig. 1.1.6 gies. Now we shall only discuss the simplest version of this concept. Let
us consider an arbitrary graph r. We choose on each of its edges an
arbitrary direction and fix it. An edge endowed with a direction (shown
Simple cycle by an arrow) will be called an oriented edge and will be denoted by ..1J,
the subscript i running through the values from 1 to O!] (see above, where
O!] is the number of edges in a graph). The same edge, but oppositely

Oriented edge aligned, will be denoted by - ..11. Each simple cycle on the graph with
the chosen direction of detour we shall write formally as an algebraic
sum of the constituent edges, taking each edge either with the "+" or
" -" sign depending on whether its direction coincides with that on the
cycle or is opposite to it. If a certain edge of the graph has not entered
in the cycle, we assume that it enters in the algebraic sum with zero co-
efficient. Consequently, each simple cycle z will be written as the sum
Cycles z = C]..1j + ... + ck..1l, where the numerical coefficients Ci are equal
to +1, -lor zero. We shall consider the various linear combinations of
such linear forms, i.e. of simple cycles written in the algebraic language.
Such general forms will not, generally speaking, correspond to simple
cycles. We shall call them cycles (without the adjective "simple"). After
Algebraic sum "grouping of like terms" such a general cycle becomes a linear form
a]..1j + ... + ak..1l, where the coefficients ai are integers already gener-
ally distinct from +I, -1, O. At the same time it is clear that the obtained
numerical coefficients aj are not quite arbitrary: they satisfy some linear
Chains relations. We might consider arbitrary linear combinations of the form
A]..11 + ... + Ak..1l, where the coefficients Ai are already arbitrary inte-
gers. Such linear combinations ar called chains. Clearly, a simple cycle
and a cycle are particular cases of chains. Not any chain is however a
cycle. A cycle is a linear combination of simple cycles.
1.1 Polyhedra 5

Now we can give an algebraic definition of the order of connected-


ness of a graph: The order of connectedness of a graph is equal to the
algebraic number of linearly independent cycles in the graph. The reader Simplex
can verify that the new definition of the order of connectedness of the
graph coincides with the geometrical definition given above.
From this example we can see that certain geometric concepts can
be transformed into algebraic concepts through an introduction of rather
simple and natural objects of algebro-geometric character (cycles, chains,
etc.). This technique proves to be very useful first of all because it al-
lows involving very powerful algebraic methods for studying geometric Barycentric coordinates
properties of figures. Employing this technique one can, in particular,
calculate qualitative topological characteristics of figures and compare
them with one another.
We now proceed to the simplest concepts of the combinatorial geo-
metry of polyhedra.

1.1.2 The Concept of an n-Dimensional Simplex. Centre of gravity


Barycentric Coordinates

Suppose in a Euclidean space ]Rn+l we are given n + 1 linearly inde- n-dimensional simplex
pendent points Ao, ... , An. We assume the points to be independent,
so there are n + 1 vectors going from the origin into these points
(Fig. 1.1.7). We place at each point Ai a non-negative m~ss mi (i.e.
assume Ai to be a material point) and require that the resultant mass of
all the points be equal to unity. This means that we have tI,.e equality
mo + ... + mn = 1. The centre of gravity of these masses (points) is the
point A, the tip of the vector OA which has the form of the linear combi-
m,
nation OA = ~Al +.. ·+mnOAn of the vectors OAi joining the origin
o with the points Ai. The numbers mo, ... , mn are called barycentric
coordinates of the point A. They are related as mi ~ 0, mo+" ·+mn = 1.
By changing the masses mo, ... , ffin we force the point A to change its
position. As a result, it runs through a certain set ,1n, which is called an
n-dimensional simplex. In other words, the simplex,1n is a convex linear
hull of the points Ao, ... ,An+l • Therefore it is occasionally referred to
as a rectilinear (or Euclidean) simplex. Figure 1.1.7 gives the simplest
examples. A zero-dimensional simplex is represented by a point, one-
6 1. Polyhedra. Simplicial Complexes. Homologies

dimensional by a segment on a plane, two-dimensional by a triangle in a


three-dimensional space. A three-dimensional simplex can be realized as
a filled tetrahedron in a three-dimensional space. The points Ao, ... , An
are called the vertices of the simplex. Clearly, a rectilinear simplex is
completely defined provided that all of its vertices are given. A simplex
is therefore denoted simply by the set of its vertices.
We shall now examine the set of points of an n-dimensional simplex
whose ith barycentric coordinate is equal to zero, i.e. mi = O. From the
definition of simplex it is immediate that this set is in turn an (n - 1)-
dimensional simplex. Besides, this simplex (which we denote by Ll~-l)
is embedded into the initial simplex Lln and is opposite to its vertex Ai
...A-
o
-+-:--......... ...'f---_I (Fig.U.8). The simplex Ll~-l is called the ith (n-l)-dimensionalface
AI. Ao AI. of the simplex Lln. Thus, an n-dimensional simplex has n + 1 faces of
Fig. 1.1.8 dimension n - 1.
Next, a k-dimensional face of the simplex Lln is the set of such
points of the n-dimensional simplex Lln for which some n-k barycentric
Simplex edges coordinates are equal to zero, while the rest k +1 barycentric coordinates
change so that the corresponding masses (coordinates) are non-negative
and their sum is equal to unity. Clearly, the k-dimensional face of the
Simplex vertices simplex Lln is itself a k-dimensional simplex.
One-dimensional faces of a simplex are usually called the simplex
edges. How many k-dimensional faces has an n-dimensional simplex?
Each face is uniquely defined by its k + 1 vertices which at the same
Face time constitute part of vertices of the initial simplex Lln. Hence, the total
number of k-dimensional faces in the simplex Lln is equal to C~!l, i.e.
to the number of permutations of n + 1 taken k + 1 at a time.
By the definition, an n-dimensional rectilinear simplex is a convex
set, i.e. a segment joining any two of its points lies entirely inside the
simplex.
The geometric boundary of an n-dimensional simplex is the union
Boundary of all of its (n - I)-dimensional faces. Each simplex has a uniquely
determined centre, i.e. a point whose barycentric coordinates are equal
to each other. This means that mo = ... = mn = I/(n + 1). Figure 1.1.8
shows the centres of one-, two- and three-dimensional simplexes.
In what follows it will be convenient for us to work with oriented
Centre of the simplex simplexes. We shall say that a simplex is oriented if a defmite order of
its vertices is given. We assume that two orders of vertices of a simplex
1.1 Polyhedra 7

differing by an even permutation determine one and the same orientation Curvilinear simp/ex
of the simplex. If the orders of vertices differ by an odd permutation,
they determine opposite orientations of the simplex. For example, if two
neighbouring vertices of a simplex exchange places (within a given or-
dering), the orientation changes (Fig. 1.1.8). We shall sometimes denote
an oriented simplex by +L1n and a simplex with opposite orientation by Positive and
- L1n. negative orientation
Here we can trace out the connection with linear algebra. The above
definition of orientation is connected with the fact that a linear mapping
of an n-dimensional simplex onto itself determined by some permutation
of its vertices has either a positive or a negative determinant depending
of whether this permutation is even or odd. So, we shall in some cases
speak of a positive or a negative simplex orientation. It should be recalled
here that any permutation of vertices of a simplex can be obtained as a
composition of elementary permutations understood as a permutation of
neighbouring vertices. At the same time, a linear mapping of a simplex
onto itself, determined by permutation of two neighbouring vertices, has
a negative determinant.
In what follows we shall consider not only rectilinear (Euclidean)
simplexes themselves, but also their various homeomorphic images, i.e.
images of a rectilinear simplex under homeomorphisms. Such homeo-
morphic images of a simplex are called topological or curvilinear sim-
plexes. Let us, for example, examine a standard tetrahedron inscribed in
a sphere, i.e. positioned inside a sphere so that all of its vertices lie on the
sphere. Then the centre of the tetrahedron coincides with the centre of
the sphere. Projecting the edges of the tetrahedron from the centre of the
sphere onto the sphere, we obtain curvilinear triangles which are the im-
ages of the equilateral faces of the tetrahedron (Fig. 1.1.9). Consequently,
the sphere falls into the union of four curvilinear triangles - homeomor-
phic images of equilateral rectilinear triangles. Figure 1.1.9 shows the
process of transformation of a rectilinear triangle into a curvilinear one.
Curvilinear one-, two- and three-dimensional simplexes are also depicted.
Thus, on each curvilinear simplex there exist vertices, curvilinear faces Fig. 1.1.9
and edges. One can compose many more objects from curvilinear sim-
plexes than from rectilinear convex simplexes. The concept of convexity
is generally not so meaningful for the curvilinear simplex as it is for the Convex simp/ex
Euclidean rectilinear one.
8 1. Polyhedra. Simplicial Complexes. Homologies

Rectilinear simplex At the same time, each curvilinear simplex "remembers" its origi-
nation from a rectilinear simplex. To this end one could fix a concrete
homeomorphism of a rectilinear simplex onto a curvilinear one. From this
Linear map it is immediately clear how the linear map of two topological curvilinear
simplexes should be naturally defined. This is a topological map under
which the pre-images of points of curvilinear simplexes (the pre-images
belonging to rectilinear simples) are carried into one another through a
linear map of the corresponding rectilinear simplexes.

1.1.3 Polyhedra. Simplicial Subdivisions of Polyhedra.


Simplicial Complexes

Using simplexes one can compose more complex objects and figures. Let
a set X of points of a Euclidean space (or, more generally, the so-called
Topological space "topological space") be represented as a union of a finite or countable
number of curvilinear simplexes of dimensions from zero to a certain
n. In other words, the set X is "glued" of curvilinear simplexes. It will
be useflli if we henceforth think of a topological (curvilinear) simplex
as a pair consisting of a rectilinear simplex and its topological map
into a certain set of points of a Euclidean space. The set of simplexes
covering X must include all the faces of these simplexes. For simplicity
we assume from now on the number of simplexes covering the set X
to be finite. We now impose simple and natural restrictions upon the
indicated subdivision.
We shall say that curvilinear simplexes form a finite simplicial sub-
Simplicial space division of the set X if the following two conditions are met.
1) There is a finite number of simplexes and each point of the set
X gets into a certain simplex (is covered by a certain simplex).
2) Either two simplexes do not intersect at all (do not have common
points), or one of them is a face of the other, or they have a cOmnlon
face which is the intersection of these simplexes.
Figure 1.1.1 0 represents all the versions of mutual disposition of two-
Fig. 1.1.10 dimensional simplexes. Figure 1.1.11 shows some "forbidden" situations.
As has already been mentioned above, with each curvilinear simplex
1.1 Polyhedra 9

there is naturally associated a corresponding topological map onto it of


a rectilinear simplex. In this case one should be aware of the condition
that two simplexes have, for instance, a common face. It is required that
under glueing together of two simplexes along their common face, into
one and the same point of the set X there go such points of faces of the
rectilinear simplexes which are associated with one another through a
linear map of one face into another. To put it differently, to the glueing
of topological simplexes (faces) there must correspond the glueing of Fig. 1.1.11
corresponding rectilinear simplexes (faces) through their linear map.
If a set of points of a Euclidean space (more generally, a topological
space) is divided into simplexes so that there hold the indicated conditions
1 and 2, this set is called a polyhedron. There exist different ways to Polyhedron
subdivide one and the same set of points into the union of simplexes
satisfying conditions 1 and 2. If one such subdivision if fixed, we shall
say that we are given a simplicial subdivision of a given polyhedron or
we are given a simplicial complex. In other words, a simplicial complex SimpliCial complex
is a set of all simplexes (and their faces) of a given subdivision of a
polyhedron, where the particular manner in which these simplexes are
glued together is indicated. One and the same polyhedron may have many
distinct simplicial subdivisions, that is, there exist many different ways Infinite polyhedra
to represent one and the same polyhedron in the form of a simplicial
complex.
In the sequel, to simplifY calculations we shall sometimes consider
divisions of a space into simplexes which are allowed to have several
common vertices (see examples for two-dimensinal polyhedra below).
Furthermore, to calculate simplicial homology groups one can employ
cells (for their exact definition see below). This means that it suffices to
represent the space as a union of closed sets the interior of each of which
is homeomorphic to an open disc (of some dimension). In particular, in
covering the space with simplexes one can reject condition 2. This does
not affect the calculation of the homology groups.
It is not at all difficult to define polyhedra glued of a countable
number of simplexes (infinite polyhedra). To do so one should require
that each point of a given set X be covered with only a finite number
of simplexes and that the neighbourhood of each point in the set X Neighbourhood
be obtained as a union of neighbourhoods of this point in each of the of the point
10 1. Polyhedra. Simplicial Complexes. Homologies

simplexes it belongs to. But further on we shall be basically interested


in finite polyhedra.
Incident Two simplexes of a simplicial complex X are called incident if one
of thttm is a face of the other. A simplicial complex (and the corre-
sponding polyhedron) is called n-dimensional if it contains at least one
n-dimensional simplex and does not contain higher-dimensional sim-
plexes.
We now sum up what has been said above.
1) A polyhedron is a set of points which can be subdivided in a
regular way into simplexes.
Simplicial complex 2) A simplicial complex is a set of simplexes and their faces which
constitute this subdivision.
3) The polyhedra defined above should naturally be called curvi-
linear or topological.
4) The set of points of a Euclidean space which is the sum of
a finite or a countable number of rectilinear simplexes satisfying the
Euclidean polyhedron conditions listed above is occasionally called a Euclidean polyhedron.
From what has been said it is clear that the topological polyhedron
can be defined in an abstract manner regardless of its embedding into
. a Euclidean space. We are not, however, interested in the extension of
generality since one can actually prove (we shall not do this here) that
Glueing any finite topological polyhedron is homeomorphic to a certain Euclidean
polyhedron i.e. can be embedded into a Euclidean space (maybe, of a
Common face large enough dimension).

1.1.4 Examples of Polyhedra

From the preceding subsection one can see that a polyhedron can be
represented as a result of glueing of a certain (finite or infinite) number
of simplexes along some of their common faces. Glueing is understood
here as follows. We take two simplexes, in each of them take one face
and identify these faces through a linear map (in case the simplexes
are rectilinear) or through the corresponding topological map (in case
the simplexes are curvilinear) (Fig. 1.1.12). This procedure gives a more
Fig.1.1.12 complex object composed of two simplexes glued along one common
1.1 Polyhedra 11

face. This provides great possibilities for visual construction of concrete


polyhedra since the process of glueing of different simplexes can be
infmitely continued.

Example 1. A Euclidean space is an infmite polyhedron. For exam-


ple, a plane can be covered (titled) with equilateral triangles. Such and
similar tilings of a plane can be seen in many gmphical canvases of
a well-known painter M.C. Escher. A three-dimensional space can be
tiled with tetrahedra (Fig. 1.1.13). In this figure one can see one of the
many ways to subdivide a standard three-dimensional cube into a union
of simplexes (tetrahedra). These tetrahedra are not equilateml. Covering
the entire space with cubes (this is easy to do) and subdividing each of
them by the indicated recipe we just obtain the subdivision of space into
simplexes, as required. A flat ring and a cylinder are homeomorphic and
are polyhedra (Fig. 1.1.13). Taking a rectangle and identifying its oppo-
site sides with orientation reversal, we obtain a Mobius strip (Fig. 1.1.13)
which is a polyhedron as well. It is natural that an individual simplex is
(the simplest) polyhedron too.

Example 2. A circle and a two-dimensional sphere are finite polyhedra.


For a circle this fact is obvious, while for a two-dimensional sphere 8 2 we
have already constructed a simplicial subdivision (see Fig. 1.1.9). It can
be readily proved, by way of a simple extension, that an n-dimensional
sphere sn is also a finite polyhedron (prove this!). Recall that the sphere Fig. 1.1.13
8n is given as a set of points in a Euclidean space ]Rn+l, such that there
holds the equation (xli+···+(xn+l)2 = r2, where xl, ... ,xn+l are
Cartesian coordinates of the point x. The number r is called the mdius
of the sphere.

Clearly, a sphere is homeomorphic to the boundary of a three-


dimensional cube. To make sure of this, it suffices to inscribe a sphere
into a cube in such a fashion that its centre coincide with the cube centre
(Fig.1.1.14) and then project the sphere from its centre onto the cube
boundary. One can readily make sure that this projection determines the
homeomorphism. Consequently, in some cases it is convenient to regard
a sphere as a two-dimensional polyhedron which is the boundary of a
standard cube. Since a square can be represented as glueing together two Fig. 1.1.14
12 1. Polyhedra. Simplicial Complexes. Homologies

triangles (Le. two-dimensional simplexes) (Fig. 1.1.14), it follows that


by subdiving all the cube faces into pairs of triangles, we obtain another
representation of a two-dimensional sphere in the form of a polyhedron.
Here the number of two-dimensional simplexes constituting the sphere
is obviously equal to 12.
It should be recalled that a sphere is the boundary of a standard
ball. We shall drill several holes in a ball, as shown in Fig. 1.1.15. As a
result, we obtain a three-dimensional set of points with a two-dimensional
boundary. This two-dimensional set appears also to admit the polyhedron
Fig.1.1.15 structure.

Example 3. Two-dimensional spheres with several "handles". We shall


begin with a particularly simple case. We shall represent a sphere in the
form of a polyhedron, as shown in Fig. 1.1.15, i.e. in the form of the plate
boundary. Not to overload the figure, we shall not depict subdivision of
side faces of the plate into triangles. We remember that any square can
be triangulated, i.e. subdivided into triangles. Then we drill a square hole
from the plate, as shown in Fig. 1.1.16. Clearly, we have obtained a new
polyhedron. Indeed, the two-dimensional boundary of a "plate with a
hole" consists of several rectangles each of which can be divided into
two triangles (by drawing a diagonal in the rectangle). Considering the
obtained two-dimensional polyhedron up to a homeomorphism, one can
easily see that it is equivalent to a "roll" (Fig. 1.1.16). This polyhedron
is called a torus (a two-dimensional torus). Applying one more home-
omorphism, one can easily transform this torus into a sphete to which
one "handle" is glued (Fig.1.1.16). So, a torus can be thought of as a
sphere with one handle. A torus can also be glued from a square, i.e. by
Fig. 1.1.16 identifying its opposite sides as indicated in Fig. 1.1.16. The first step is
to glue two sides b to obtain a cylinder. The second step is to glue the
bases of the cylinder to obtain a torus.
r------ -----------
fB
r

Now let us examine a plate with two holes drilled (Fig. 1.1.17).
I
" ~
rr
____ ..,r Clearly, we again obtain a two-dimensional polyhedron composed of two
I
I rectangles each of which can be subdivided into simplexes. Applying an
appropriate homeomorphism, one can transform a ''plate with two holes"
into the object depicted in Fig. 1.1.17 and called in the mathematical

Fig.1.1.17
1.1 Polyhedra 13

slang a ''pretzel''. If we go on transforming this figure, we can represent


it in the form of a sphere with two handles, as shown in Fig. 1.1.17.
The described process of drilling holes in a plate can be continued. It
is clear that in a plate one can drill an arbitrary finite number of square
holes (Fig. 1.1.18). Applying appropriate homeomorphisms, we obtain
spheres with an arbitrary number of handles. Hence, we have discovered
an infinite series of two-dimensional polyhedra - a sphere with 9 handles.
The number 9 is called the genus of a surface.
As has already been mentioned above, we shall allow some sim-
plexes to have two common vertices. This will not affect our calculations
of homology groups and will simplify significantly (in some cases) the
representation of topological spaces as unions of simplexes.

Example 4. A projective plane. Recall that polyhedra can be obtained


by means of glueing simplexes along their common faces. Figure 1.1.19 Fig. 1.1.18
gives an example of a polyhedron obtained by glueing eight triangles
which compose the subdivision of the square. Those sides of the trian-
gles which should be glued together are labelled with identical letters, the
directions of the arrows showing the orientation for glueing correspond-
ing sides (segments). Ifwe digress from the system of triangles and view J.
the whole square, it becomes obvious that the polyhedron in question will
be obtained if on the sides of the square we identify the pair a and a c/
J..~.::.:..
. a ~cJ
(with allowance made for the arrows) and the pair (3 and (3 (also with .1 d., J
:; ••::: ••• :.:.

allowance made for the arrows) (Fig. 1.1.19). The polyhedron obtained "" ::.....::... :-:
::. c I
a
is called a projective plane or a two-dimensional projective space. In the
sequel we shall repeatedly encounter this polyhedron and get acquainted Fig.1.1.19
in more detail with its various interesting properties.

Example 5. A Klein bottle. We again take a square divided into eight


triangles, but change (as compared to Example 4) the rule of glueing
the sides of the square according to Fig. 1.1.20. The polyhedron thus
obtained is called a Klein bottle. It is convenient to represent it in a three-
dimensional space as follows. First we glue two sides (3 and (3 to obtain
a cylinder (Fig. 1.1.20) on whose boundary circles we see the arrows
a and a. To identify these circles (so that the directions of the arrows
coincide), one should, unfortunately, "puncture" the surface, which leads
to self-intersection.
Fig. 1.1.20
14 1. Polyhedra. Simplicial Complexes. Homologies

Among the variety of all polyhedra there exists an important class


Topological manifolds of polyhedra called manifolds (topological manifolds). A polyhedron of
dimension n is called an n-dimensional manifold (or an n-manifold) if
each of its points has a neighbourhood (inside the polyhedron) which
can be homeomorphically mapped onto the interior of a standard n-
dimensional ball Dn. Recall that a standard open ball is the set of points
x in a Euclidean space ]Rn, satisfying the inequality (xl )2+ .. .+(xn)2 < 1.
A closed ball (or a ball with boundary) is the set of points x satisfying
Bing house the inequality (xl)2 + ... (xni :::; 1. Clearly, a closed ball is obtained

+M
from an open one by adding a boundary sphere.
The examples of polyhedra listed above (Euclidean space, sphere,
sphere with handles, projective plane, Klein bottle) are in fact manifolds.
Figure 1.1.21 demonstrates two polyhedra (one-dimensional and two-
Fig. 1.1.21 dimensional) which are not manifolds. In both cases the point 0 does not
have a neighbourhood (in the polyhedron) homeomorphic, respectively,
to a one-dimensional disc (i.e. to an interval) and a two-dimensional disc.

Example 6. "Bing house". This two-dimensional polyhedron is needed


for dealing with the problems of three-dimensional topology discussed
in chapters to follow. To construct this polyhedron, we take a cube and
divide it into two parts by the equatorial horizontal plane (Fig. 1.1.22).
On the upper face of the cube we start "eating out" a vertical hole
downwards. We reach the equatorial plane, "eat it away" to emerge in
the lower chamber and "eat out" the whole of it except for the figure B
, which has been formed by a similar construction starting with the lower
, ,,
,, face of the cube. Analogously, we "cut our way" into the cube from
, below, "eat out" a hole inside the B, puncture the equatorial plane, run
I

,,
I

Fig. into the upper chamber and "eat out" its interior, except for the walls of
I

1.1.22 the construction A.

1.1.5 Barycentric Subdivision

Open ball As has already been mentioned, there exist many ways to represent each
polyhedron in the form of a simplicial complex, i.e. to divide it into
Closed ball simplexes. Such ways are infinitely many. To have a deeper insight into
1.1 Polyhedra 15

polyhedra, it turns out useful to distinguish a special class of simplicial


divisions called barycentric subdivisions (of a given simplicial division). Barycentric subdivision
Generally speaking, the operation of subdivision (reduction) of a sim-
plicial complex consists in the following. We divide each simplex into
smaller simplexes in such a way that we again come to a simplicial com-
plex. Clearly, the polyhedron remains unchanged under this operation. Centre of a
Affected will be the simplicial complex whose support (''body'') is a curvilinear simp/ex
given polyhedron. In the sequel it is good practice to view a polyhedron
as a set of points and a simplicial complex as a certain "scheme" of
this set which results from representing the set as a union of simplexes.
And again we can reduce the simplex to fragments in many ways. We
are particularly interested in one special way of this reduction. We now
proceed to its exact description.
We consider all the vertices of a given simplicial complex. We shall
not change them (a point cannot be reduced to smaller fragments). Then
examine all one-dimensional edges, i.e. one-dimensional simplexes. We
introduce (add) new vertices taking the centres of the one-dimensional
simplexes. Each edge splits into two one-dimensional simplexes. Then
we take all the two-dimensional simplexes and the new vertices will
be the centres of these simplexes. We assume the concept of the cen-
tre of a curvilinear simplex be well defined since all our simplexes are
always topological images of rectilinear simplexes whose centre (the cen-
tre of gravity) is uniquely defmed (Fig. 1.1.23). From the centre of the Fig. 1.1.23
two-dimensional simplex we project (using straight lines on the rec-
tilinear simplex) already subdivided one-dimensional boundaries (i.e.
edges). In other words, we draw medians in the triangle. They inter-
sect in the centre of gravity of the triangle (Fig. 1.1.23). Hence, each
one-dimensional simplex has fallen into two one-dimensional simplexes.
Each two-dimensional simplex has fallen into six two-dimensional sim-
plexes. Now take three-dimensional simplexes. In each of them we es-
tablish the centre and project from it already subdivided two-dimensional
faces of this simplex. As a result, we divide the three-dimensional sim-
plex into 6 . 4 = 24 smaller three-dimensional simplexes. Clearly, this Centre of gravity
process can be extended to higher dimensions. As a result, we are led
to a new simplicial complex which is called a barycentric subdivision of
the initial one.
16 1. Polyhedra. Simplicial Complexes. Homologies

The process of barycentric subdivision can be repeated, i.e. one can


Process of realize a two-, three- and m-stage barycentric subdivisions. It is intu-
barycentric subdivision itively clear that by increasing m we shall be able to divide the initial
simplicial complex into simplexes of arbitrarily small size. So, reduc-
ing the initial simplicial division into smaller fractions, we cover the
initial polyhedron with increasingly small simplexes whose sizes are si-
multaneously and uniformly decrease. Visually, this process leads to the
following. If a polyhedron is divided into a small number of simplexes,
then it is a sufficiently rigid body provided that each simplex is regarded
as a rigid body and that simplexes are allowed to rotate about their com-
mon faces. Increasing the number of simplexes, we introduce a larger
number of edges about which the neighbouring simplexes can rotate. As
a result, the polyhedron becomes "softer", pliable and deformable by a
larger number of ways.
Suppose 0 is a vertex of a simplicial complex. Consider all the
simplexes incident to this vertex. Their union is called the simplicial star
of the given vertex. Clearly, it is the neighbourhood of the given vertex
Fig. 1.1.24 in the polyhedron (Fig.1.1.24).

1.1.6 Visual Material

Simplicial star In this section we provide a visual demonstration of polyhedra as an


of a vertex example of geometric constructions. Figure 1.1.25 gives an example of
a rather complicated polyhedron, partitioned into a union of cubes and
parallelepipeds. As is known, all these objects can in turn be trianglated,
which will just give us triangulation of polyhedra. But it turns out that one
can consider cubic partitions of polyhedra, i.e. take for the elementary
object a cube instead of a simplex. This theory has both advantages and
disadvantages. On the basis of this theory one can determine the so-called
Cubic homologies cubic homologies (compare with the material of Sect. 1.2).
A similar idea (but in a more rectilinear version) is expressed in
Cubic partitions Fig. 1.1.26. For the sake of visualization, it is generally convenient to
think of polyhedra (even if they are smooth manifolds) as of constructions
composed (glued) of cubes, parallelepipeds. This image is apparently
psychologically more convenient than the one composed of tetrahedra.
Cubic polyhedron Figure 1.1.26 illustrates a "cubic polyhedron" with a cut. This image

Page 17, Fig. 1.1.25. Simplicial polyhedra. Cubic partitions and cubic ho-
mologies. To M. Bulgakov's "The Master and Margaret": night talk be-
tween Pontius Pilate and Afranii (story of Judas)
18 1. Polyhedra. Simplicial Complexes. Homologies

will soon be of use in the discussion of the concept of chains and of


the important algebro-geometric fact that summation of chains leads in
some cases to mutual annihilation of common pieces of the boundary.
"Glueing" the crack in the polyhedron in Fig. 1.1.26, we thus obviously
decrease its boundary: the crack boundaries are mutually annihilated.
Here the "crack" separates the simplicial chain (polyhedron) into two
parts with "large boundaries", but when glueing the crack, we decrease
Common boundary the common boundary of the chain, i.e. 8(A+B) =8A+8B-(8An8B).
Figure 1.1.27 represents prisms into the union of which there splits the
Boundary of the chain direct product of the polyhedron by a segment.
Some of the mathematical figures included in the book have in addi-
tion to their primary mathematical meaning also another, "non-formal",
extra-mathematical, associative meaning. For example, the drawings by
the author, listed above, were once intended not only as mathematical
illustrations, but also as illustrations to the remarkable novel by M.A.
M.A. Bulgakov Bulgakov "Master and Margaret". About 40 illustrations were created by
the authors many years ago for the first Russian edition of Bulgakov's
novel. But this book was published without any pictures. Now the reader
of the present mathematical book can see not only the geometrical con-
tent of these drawings but also can try to connect them with remarkable
story about Jesus Christ and Pontius Pilate (from Bulgakov's novel):
Christ is delivered to Pilate, story of Judas, cruciftxion of Christ, story
of Matthew. These are Figs. 3.3.16; 3.1.12; 1.3.13; 1.3.14; 2.5.39; 1.3.10;
4.1.6; 2.5.37.

1.2 Simplicial Homology Groups of Simplicial Complexes


(Polyhedra)

1.2.1 Simplicial Chains

Simplicial chains The language of polyhedra and simplicial homology groups (see below)
is rather obvious and convenient for the first acquaintance with the im-
Simplicial portant geometric concepts. This language is primary enough, based on
homology groups

Page 19, Fig. 1.1.26. Cubic polyhedron and the action of boundary oper-
ator. To M. Bulgakov's "The Master and Margaret"; crucifixion of Jesus
Christ and hurakan over Jerusalem

Page 20, Fig. 1.1.27. Simplicial prisms and cubic homology groups. To M.
Bulgakov's "The Master and Margaret"; crucifixion of Jesus Christ; story
of St. Matthew - the Jesus' pupil; Gospel according to St. Matthew
1.2 Simplicial Homology Groups of Simplicial Complexes (polyhedra) 21

a small number of concepts, although in concrete calculations it is some-


times cumbersome. In modem geometry, the calculation of homology
groups is usually carried out in other, more versatile terms, for instance
in terms of the so-called cell homologies. But the relative simplicity of Cell homologies
calculation of cell homologies rests on a fairly sophisticated procedures
of the very definition of cell homology groups. The advantages and disad- Cell homology groups
vantages of the different definitions of homology groups can be tabulated
as follows.
Simplicial The definition is The proof of Concrete calcu-
homolo- simple the invariance lations are rather
gies is complicated complicates
Cell The definition is The proof of Concrete calcu-
homolo- rather complica- the invariance lations are much
gies ted of simplicial is simple simpler than
homologies those

This shows the advantage of cell homologies since in concrete in-


vestigations it is preferable to deal with easily computable (even though
complicatedly defmed) objects. We have however preferred to acquaint
the reader with simplicial homologies (and, accordingly, with polyhedra)
since our prime concern is visual and elementary character of the mate-
rial. The acquaintance with simplicial homologies makes it possible to
comprehend very quickly the very idea of homologies which plays an Idea of homologies
important role in modem geometry, mechanics and physics. The same
general idea underlies the concept of cell homologies, and therefore we
think it fruitful to apply to the language of polyhedra and simplicial com-
plexes at early stages of studying the fundamentals of topology. After
this it is easy to go over to cell homologies.
We shall consider a polyhedron X and fix its arbitrary simplicial
division. For simplicity, we shall further on denote the corresponding
simplicial complex by the same letter X. We shall examine the set of
all k-dimensional simplexes of the polyhedron X, number them in some
(arbitrary) order and associate with each simplex (also in an arbitrary
way) some orientation. We denote these simplexes by Llf, where i is
the ordinal number and k is the dimension of the simplex. The number
i can change from unity to infmity. Simplex numbering and orientation
are assumed to be fixed. Let us consider some Abelian group G, for Abelian group
example, a group of integer number Z, a group of real numbers I. or
22 1. Polyhedra. Simplicial Complexes. Homologies

a finite group Zp of residues modulo p. The indicated groups are most


frequently used in topology. For definiteness, we shall for the present be
concerned with the case G = Z. An extension of all further constructions
to the case of an arbitrary Abelian group is made automatically and does
not add any new essential points.
We shall examine linear combinations of the form c = 2:i aiLl7,
where ai are integers (either positive or negative) and ..17 are k-
dimensional simplexes, and assume the sum to contain only a finite
Integer-valued chains number of nonzero ai. Such linear combinations are called integer val-
ued k-dimensional chains (or simplicial integer-valued k-dimensional
Integer-valued function chains). Such a chain can be interpreted as an integer-valued function
defined on the set of all k-dimensional simplexes, i.e. c: ..17 --t ai for
all i. This function is assumed to be nonzero only on a finite number
of simplexes and to be odd, i.e. c ( - ..17) = -ai = -c (..17), where by
-..17 we denote the simplex ..17 with a reverse orientation. Working with
chains we shall henceforth mainly use the language of linear forms for
it is convenient for calculations.
The simplest examples of chains are those of the form 1 . ..17 and
Elementary chains ( -1)· ..17 which are sometimes called elementary. Chains can be summed
up as ordinary linear forms. Namely, the sum Cl + C2 of two chains
Cl = 2:i aiLl7 and C2 = 2:i biLl7 is the chain Cl + C2 = 2:/ai + bi ) ..17·
Group of simplicial Consequently, the set of all k-dimensional integer-valued chains forms
integer-valued chains an Abelian group which we denote by Ck(X). The generators of this
group are elementary chains of the form 1 . ..17.

Definition. The group Ck(X) is called the group ofk-dimensional simpli-


cial integer-valued chains ofthe polyhedron X (ofthe simplicial complex
X.)

This shows how one can define chains with coefficients from an
arbitrary Abelian group. To this end one should consider linear forms
7,
2:i gi Ll where the coefficients gi belong to the abelian group G. The
arising Abelian group of chains with coefficients in the group G will be
denoted by Ck(X, G). In this notation, the above-defined group Ck(X) of
integer-valued chains will be written as Ck(X, Z). This is afree Abelian
Free Abelian group group because it is represented in the form of the direct sum of a certain
number of infinite Abelian groups Z.
1.2 Simplicial Homology Groups of Simplicial Complexes (Polyhedra) 23

If the polyhedron X is finite, the group of its k-dimensional integer-


valued chains has ajinite number ofgenerators (i.e. is a finitely generated Finite number
Abelian group). of generators
Thus, we have associated with each polyhedron X (each simplicial
complex X) a set of Abelian groups Co(X), CI (X), ... , Cn(X). Clearly,
for k < 0 and k > n the groups of chains are undefined since there are no
simplexes with dimension k less than zero or greater than n = dim X.
However, it is sometimes convenient to assume (by definition) all the
groups Ck(X), where k > n, to be zero.

1.2.2 Chain Boundary Chain boundary

We shall consider an integer-valued elementary chain 1 . Llk (which


we denote for simplicity by Llk) and defme its algebraic boundary as Algebraic boundary
a certain (k - 1)-dimensional integer-valued chain which we shall now
represent by an explicit formula. As a preliminary step we shall discuss
the concept of orientation induced on the ith face Ll~-I of the simplex Induced orientation
Llk. Recall that the orientation of the simplex Llk is assumed to be
given and fixed. The simplex Llk is given by the set of its vertices
Ao, AI'"'' Ak (see Sect. 1.1). Clearly, the (k - 1)-dimensional faces
Ll~-I of the simplex Ll k are obtained by a successive elimination of its One-dimensional simplex
vertices, i.e. Ll~-I = (Ao, ... ,Ai-I,Ai+l, ... ,Ak). We assume that the
orientation of the face Ll~-I (with the number i) induced on it by the
orientation of the simplex Llk is determined by the sign (_1)i.
. .... . ...
-zEi
A. ,11. At
We now give several examples. A one-dimensional simplex (a seg- ~

ment) Lli is given by its two vertices Ao and Al (Fig.1.2.l). Then its A1 +
faces are represented by a pair of points Ao and Al taken one (Ao) with
6 -+
+: f).:.~
a "-" and the other (AI) with the "+" sign. Indeed, Lli = (Ao, AI), ..........
Ll~ = (-1)0 Al = AI; Ll~ = (_1)1 Ao = -Ao. Geometrically, the differ- Ao - A:z.
ence in the sign is explained by the fact that at the point Ao (Fig.1.2.l) Fig. 1.2.1
the arrow indicating the simplex orientation is turned inside the simplex,
while at the point Al outside the simplex. So, when moving along the
straight line on which the segment lies, we first enter the segment (the
"-" sign) and then leave it (the "+" sign). Therefore, the concept offace
orientation introduced above corresponds to the intuitive idea. We shall
consider the following example.
24 1. Polyhedra. Simplicial Complexes. Homologies

Two-dimensional simplex A two-dimensional simplex is given by three vertices Llz = (Ao, AI,
Az) (Fig. 1.2.1). The simplex has three one-dimensional faces (sides
of the triangle), namely, an edge Ll~ = (-I)o(AI,A z) = (AI, Az),
an edge .11 = (-1)1 (Ao,Az) = -(Ao,Az) = (A2,Ao), and an edge
.11 = (-Ii (Ao,AI) = (Ao,A I ). In Fig. 1.2.1 the arrows are placed on
one-dimensional faces of the simplex to show the orientation induced
on these faces. It should be emphasized that the edges (AI, A2) and
(A o, AI) are positively oriented and the edge (Ao, A2) is negatively ori-
Boundary edges ented. Therefore, we have replaced the edge -(Ao, A2) by an equivalent
edge (A2, Ao). From Fig. 1.2.1 it is seen that the formal definition of
induced orientation, given above, is in close agreement with geomet-
rical intuition. Indeed, from the calculations we have revealed that the
boundary edges endowed with induced orientation give one and the same
direction ofrotation on the triangle. Going round its boundary clockwise,
it would be natural to expect that the passed-by edges are endowed with
arrows indicating the same direction of motion. This is precisely what we
Algebraic boundary have obtained on the basis of the algebraic definition of induced orien-
of a simplex tation. Now we can define an important concept of algebraic boundary
of a simplex.

Induced orientation Definition. The boundary 8Llk of an oriented simplex Llk is the sum of
all its (k - I)-dimensional faces taken with induced orientation.

Boundary operator We shall write the simplex boundary in the algebraic language. We
obtain:
k
aLlk = L(_1)i Ll~-I = Ll~-I - Ll~-I + Ll~-I - ... Ll~-I
i=O

In terms of vertices, this formula becomes


k
8(Ao, ... , Ak) = L( _I)i (Ao, ... , Ai-J, ... , Ai-I, ... , Ak)
i=O
In other words, we eliminate successively the simplex vertices, obtain the
corresponding faces (of dimension smaller by unity) and sum up (in the
form of a formal linear combination) taking them with the sign (_I)i,
where i is the number of the eliminated vertex. It should be clarified
1.2 Simplicial Homology Groups of Simplicial Complexes (Polyhedra) 25

that the (k - I)-dimensional simplex ..1~-1 obtained after the vertex Ai


has been eliminated is depicted geometrically as the face opposite (in the Opposite face
simplex ..1k) to the vertex Ai (Fig. 1.2.2).
It is now obvious that this algebraic definition of the simplex boun-
dary (the algebraic character manifests itself in the allowance made for
the sign, i.e. for orientation) is quite natural. Indeed, the boundary of a
one-dimensional simplex is the sum of its vertices of the form:

The boundary of a two-dimensional simplex is the sum of its edges:

All these three edges determine one and the same clockwise rotation,
that is, precisely that orientation which is induced on the boundary of a
triangle if inside it one sets a clockwise rotation.
Weare now in a position to derme the boundary of an arbitrary Boundary of a chain
chain. Suppose e = al..1f + ... + aq..1~ is a simplicial integer-valued
chain.

Definition. The boundary of a k-dimensional chain e is a (k - 1)-


dimensional integer-valued simplicial chain 8e given by the following
explicit formula:

Given this, we assume that the meaning of the expression 8..1~


has been defined above. In other words, on setting the operator 8 on Boundary operator
elementary chains 1 . ..1j, we then extend it "by linearity" to arbitrary
linear combinations of elementary chains thus obtaining a well dermed
operator on the entire Abelian group of chains Ck(X). This operator is
called the boundary operator, or the operator of the boundary, or the
operator of taking the boundary.
26 1. Polyhedra. Simplicial Complexes. Homologies

1.2.3 The Simplest Properties of the Boundary Operator. Cycles.


Boundaries

a
1) The operator is linear. This means that for any k-dimensional
chains CI, C2 and for any integer coefficients a, b there holds the identity
o(aci + bC2) = aoC) + bOC2. To say it differently, the operator a deter-
Abelian group mines a homeomorphism of the Abelian group of k-dimensional chains
of k-dimensional chains Ck(X) into the Abelian (k - I)-dimensional chain group Ck-I (X). We
shall sometimes denote this operator by Ok indicating explicitly the di-
mensions of the chains on which the operator is defined. Consequently,
in the preceding subsection we have in fact defined the whole family of
operators 80, 01, ... , Ok, ... , an, where n = dim X (Le. the dimension of
Square of the the complex X). Thus, Ok: Ck(X) -+ Ck-I (X).
boundary operator 2) The square of the boundary operator ais identically zero.

a
Proof. Since the operator is linear, it suffices to prove the statement
only for elementary chains, Le. for a single k-dimensional simplex ..1k .
It is sometimes convenient to write the boundary operator in the form
O..1 k = Ef=o(-1 )i(Ao, ... , Ai, ... ,Ak),where the sign" (the hat) im-
A "

plies that the symbol under it is omitted, eliminated. We must prove that
{j2 == 0 a
02..1 k = O. Applying the operator for the second time, we obtain a
long sum which necessarily involves the following two summands:

(-I)i(-IY (Ao, ... ,Aj,'" ,Ai'''' ,Ak)


Formula for
and
boundary operator . . 1 A A

(-1)1(-1)'- (A o, ... , Aj , ... , Ai, ... , Ak )


We can assume without loss of generality that j < i. In the first summand
the vertex Ai was discarded with the first application of the operator a,
Le. before the vertex Aj was discarded (already with the second applica-
tion of the operator a). In the second summand the picture is different.
Here the vertex Aj was first to be discarded (with the first application of
a), and only after that it was the vertex Ai with the second application of
a). Consequently, to know the sign when we discard the vertex Ai, we
should count the number of vertices leftward of Ai. It is clear that left-
ward of it there are i-I vertices, for the vertex Aj was discarded in the
preceding step. Since the signs (-I)i-J and (-1 )i+j-l are opposite, the
1.2 Simplicial Homology Groups of Simplicial Complexes (Polyhedra) 27

two summands that we have distinguished enter in the overall sum with
opposite sign and therefore are mutually annihilated, which completes
the proof.

Its geometrical meaning is absolutely clear: the boundary of a simplex Boundary of a simplex
has no boundary. has no boundary
3) Ijo(ae) = 0 and at-O, then oe= O. This assertion is obvious, as
we are examining integer-valued chains and from the equality o(ae) = 0,
where at-O, it follows that aoe = 0, i.e. a(oe) = 0 and oe = 0 (the
coefficient at- 0 can be cancelled out).

Definition. A chain z is called a cycle if its boundary is equal to zero, Cycle


i.e. OZ = O. A chain b is called a boundary if it is representable as b = oh,
i.e. in the form of the boundary of a certain chain h whose dimension is Boundary
greater by unity.

Clearly, all these objects (boundary operator, cycles, boundaries) can also
be defined for an arbitrary Abelian group of coefficients. We shall not
repeat here the corresponding definitions and invite the reader to do it
himself.
4) The set of cycles forms an Abelian subgroup in a chain group.
This subgroup is denoted by Zk(X). The set of boundaries forms an Set of boundaries
Abelian subgroup in a chain group. This subgroup is denoted'by Bk(X).
Each boundary is a cycle. This means that the subgroup Bk(X) is always Each boundary
contained in the subgroup Zk(X). These statements follow from linearity is a cycle
of the operator 0 and from the equality & == 0 proved above. A cycle
should not necessarily be a boundary. This means that in the general
case the group Zk(X) can be larger than the subgroup Bk(X). So, Ok:
Ck(X) --+ Bk_I(X); Ok: Zk(X) --+ 0, Ok-I: Bk_I(X) --+ O.

1.2.4 Examples of Calculations of the Boundary Operator

We shall assume (by definition) that the boundary ofany zero-dimensional


chain is equal to zero (since there exist no nonzero chains of negative
dimension). So long as any zero-dimensional chains a linear combina-
tion of points (vertices of a polyhedron) with integer coefficients, our
28 1. Polyhedra. Simplicial Complexes. Homologies

agreement is demonstrative: a point has no boundary (i.e. the boundary


Any zero-dimensional of a point is equal to zero). Accordingly, any zero-dimensional chain is
chain is a cycle a cycle.
We shall consider a two-dimensional simplex ..12 as an elementary
chain. Then its boundary 8..12 has the form Cl + C2 + C3 (Fig. 1.2.3).
Therefore, the one-dimensional chain c =Cl +C2 +C3 is the boundary. At
the same time, it is obviously a one-dimensional cycle.
We shall consider a square divided into four two-dimensional sim-
plexes (Fig. 1.2.3) on each of which we determine a clockwise rotation.
We shall take the chain which is the sum of these four simplexes with
coeffi(;ients 1 and calculate its boundary. To do so, we should calculate
the boundaries of all the four simplexes and then sum up these boun-
daries. Given this, the following will happen. Each one-dimensional edge
lying inside a square will enter this sum twice, each time with a different
sign. The point is that orientations induced on an edge by two adjoining
triangles are reverse. Consequently, all the edges p, q, T, 5 will cancel
out and will not enter the final sum. At the same time, each of the one-
dimensional edge a, b, c, d will enter the sum only once, and therefore
the boundary of the square regarded as a two-dimensional chain will co-
incide with its geometric boundary - the sum of the four oriented edges
a+b+c+d.
Let us consider a square with a hole (Fig. 1.2.3) triangulated into
eight triangles. We endow each of them with orientation (clockwise ro-
tation) and take a two-dimensional chain c which is the sum of all these
Fig. 1.2.3 triangles with coefficients 1. Acting as in the preceding case we see that
all the interior one-dimensional edges are mutually annihilated. As a re-
sult, the boundary of the chain will consist of two sumes: a + b + c + d
(with a "+" sign) and m+n+h+l (with a "- sign). Thus, the algebraic
Clockwise rotation boundary of a square with a hole coincides with its geometric boundary,
the exterior part of the boundary being taken with a "+" and the interior
with a "-" sign. Consequently, 8c =a+b+c+d - m - n - h -t. Since
a square with a hole is homeomorphic to a ring, the algebraic boundary
of a ring consists of an exterior circle (with a "+" sign) and an interior
circle (with a "-" sign.)
From the demonstrative point of view, a cycle may therefore be
Closed surface thought of as a "closed surface", i.e. a surface without boundary. The
1.2 Simplicial Homology Groups of Simplicial Complexes (polyhedra) 29

boundary ofa chain is its geometric boundary if all the chain-constituting


simplexes are taken with coefficient 1 (Fig. 1.2.4).
In the preceding subsection we have mentioned that a cycle is not
necessarily a boundary. For example, if we triangulate a two-dimensional
sphere and take the chain which is the sum of all the triangles obtained
with coefficients 1, we obviously come to a two-dimensional cycle. At
the same time, it cannot be represented as the boundary of a certain
three-dimensional chain for the simple reason that our complex does not
have a single nonzero three-dimensional chain (the sphere dimension is
equal to two). We could also give more delicate examples (we shall soon Fig. 1.2.4
present them) when cycles are not boundaries.

1.2.5 Simplicial Homology Groups

Let Z be a k-dimensional cycle, i.e. 8z = O. We shall say that it is Geometric boundary


homologic to zero if it is the boundary, i.e. z = 8h for a certain (k + 1)-
dimensional chain h. By the way, in this case the chain h is generally
speaking not uniquely defined. If z = 8h for some chain h, then there also
holds the equality z = 8(h+I), where l is an arbitrary (k+ I)-dimensional
cycle, i.e. 8l = O.
Two k-dimensional cycles z\ and Z2 will be called homologic, if Homologic cycles
their difference z\ - Z2 is homologic to zero, i.e. z\ - Z2 = 8h for a
(k + I)-dimensional chain h. The fact that the cycle z is homologic to
zero is sometimes written as z rv O. Then the homology of the two cycles
z\ and Z2 is written as Zl - Z2 rv 0 or Zl rv Z2.
The relation z\ rv Z2 is occasionally called homology. If the cycle Homology
Zl is homologous to the cycle Z2, they differ by a boundary, i.e. there
holds the equality Z\ = Z2 + 8h.
The homology relation is determined not only for cycles, but for
chains as well. The chains Cl and C2 are called homologous (ct rv C2) if Homologous chains
they differ by a certain boundary, i.e. C\ = C2 + 8h or ct - C2 rv O.
We now return to cycles. Suppose X is an arbitrary polyhedron,
Zk(X) is its k-dimensional cycle group, Bk(X) the group of its k-
dimensional boundaries. Since the group Bk(X) is always contained in Group of boundaries
the group Zk(X) and both the groups are Abelian, it follows that the
factor quotient group Hk(X) = Zk(X)/ Bk(X) is well defined.
30 1. Polyhedra. Simplicial Complexes. Homologies

Definition. The group Hk(X) is called the k-dimensional integer-valued


simplicial homology group of a polyhedron X. It is an Abelian group.

If we consider chains with coefficients in an arbitrary Abelian group


G, then, repeating almost literally all the previous constructions, we
would arrive at the definition of the homology groups of the polyhe-
dron X with coefficients in the group G. These groups are denoted by
Hk(X, G), i.e. the group of coefficients is indicated In this notation, the
Integral-valued simplicial integer-valued homology groups Hk(X) will be written as Hk(X, Z). We
homology group shall use both depending on the context.
We shall discuss the concept of homology group in terms of ho-
mologous cycles. Let us consider the group of cycles Zk(X). Any linear
Homology groups combination of cycles homologous to zero is again a cycle homologous
with coefficients to zero. This just means that Bk(X) is a subgroup in Zk(X). Therefore,
in the group G for each cycle z one can consider its coset with respect to the subgroup
Bk(X), i.e. a set of cycles of the form z + oh, where h runs through all
(k+ I)-dimensional chains. The coset obtained can be denoted by {z}.
Then it is clear that the two classes {zI} and {Z2} either coincide or do
Group of coefficients not intersect. Clearly, the linear combination of the classes {Zl} and {Z2}
is well defined, namely: a{zl}+b{Z2} = {azl +bZ2}. Thus, the group
Hk(X) is interpreted as a group of classes of the form {z}. However,
when speaking of elements o~ the group Hk(X) we shall often bear in
mind the cycles themselves, i.e. some representatives of the class {z}.
We shall point out an essential property of the homology group
Hk(X). From the fact that the cycle mz is homologous to zero (where
m is an integer) it does not at all follow that in the general case the cycle z
itself is homologous to zero. In other words, "homologies are indivisible".
From the point of view of the group Hk(X) this means that there may
exist nonzero elements {z} from the group Hk(X), such that their integer
multiple m {z} is equal to zero (in the group Hk ), i.e. mz '" 0, although
Finite-order elements z rf O. Such elements are therefore finite-order elements in the group
Hk(X). This implies that the homology groups Hk are not generally free
Abelian groups. They may contain subgroups consisting of finite-order
Homologies elements. Thus, each group Hk can be represented as the direct sum of its
are indivisible two subgroups Ak and Bk, where Ak is a free Abelian group (the direct
sum of a certain number of copies of the group Z) and Bk is a finite
Abelian group. The group Ak is uniquely characterized by the number
1.2 Simplicial Homology Groups of Simplicial Complexes (Polyhedra) 31

(3k of constituent copies of the group Z. In the algebraic language, the


number (3k is the rank of the group Ak, i.e. the minimal number of Rank of the group
genemtors of this group. At the same time it is clear that the number (3k
is also the mnk of the entire group Hk, i.e. (in this case) the minimal
number of free generators (infinite-order genemtors).

Definition. The number (3k is called a k-dimensional Betti number of a k-dimensional


polyhedron X. Betti number

To summarize we shall say that to each n-dimensional polyhedron X


we have assigned a set of Abelian groups Ho(X), ... , Hn(X) (homology
groups). These groups are Abelian. The homology groups Hk(X) are Chain complex
equal to zero in all the dimensions k exceeding n (the dimension of the
polyhedron). These groups have been calculated for a certain simplicial
polyhedron partition. If the simplicial partition is fixed then associated
with it is the following sequence of free Abelian groups Ck(X) (chain
groups) and homomorphisms Ok (boundary operators): 0 ~ Co(X) ~
C1(X) ~ ... ~ Ck(X) ~ ... ~ Cn(X) ~ O. This sequence is
called a chain complex (of a given simplicial complex).
As we know, however, one can represent one and the same polyhe-
dron in different ways in the form of a simplicial complex, i.e. triangu-
late it differently. But to different simplicial complexes there genemlly
correspond different chain complexes. The simplicial homology groups In variance of the
Hk(X) might therefore be expected to depend on the choice of poly- simplicial homology groups
hedron triangulation. But this is not the case. The simplicial homology
groups appear to be determined only by the polyhedron itself but not by
its triangulation.

Theorem. The simplicial homology groups of the polyhedron do not Different triangulation
depend on the way the polyhedron is represented as a simplicial complex. of the polyhedron

.This fact increases sharply the importance of homology groups trans-


forming them into algebmic objects that characterize the polyhedron it-
self and not the way of its determination (representation). In what follows
we shall mainly deal with finite polyhedra. In this case all the homology
groups Hk(X) are finitely generated. Each group Hk(X) is represented
32 1. Polyhedra. Simplicial Complexes. Homologies

in the form of the direct sum

Hk(X) = Z EEl .•. EEl Z EElZPI EEl ••• EEl Zp, ,


'---..-'
(3k

Betti number where 13k is the rank (the Betti number), ZPi are fmite cyclic Abelian
groups of orders Pi, and each number Pi may be assumed to be a divisor
of a preceding one. The numbers PI , ... , Ps are sometimes referred to as
Torsion coefficients torsion coefficients of the simplicial complex X in dimension k.

1.2.6 Examples of Calculations of Homology Groups. Homologies


of Two-dimensional Surfaces

Zero-dimensional 1) Zero-dimensional homology groups. To begin with, we shall


homology groups calculate a zero-dimensional homology group Ho(X) of a linearly (ar-
cwise) connected polyhedron X. A polyhedron (a simplicial complex)
Linearly is called linearly connected if any pair of its points can be joined by a
connected polyhedron continuous arc lying wholly in the polyhedron. We state that the zero-
dimensional homology group of a linearly connected simplicial complex
is isomorphic to a group Z. If the simplicial complex is disconnected,
Any two vertices of its zero-dimensional homology group is isomorphic to the direct sum
a connected polyhedron Z EEl ••. EEl Z = zq, where the number of summands q equal to the number
are homologous of components of linear connectedness of the complex X.
As we already know, the zero-dimensional cycle group coincides
with the zero-dimensional chain group, i.e. Co(X) = Zo(X). A zero-
dimensional cycle is a linear combination of vertices of a simplicial
complex with arbitrary integer coefficients. We fix an arbitrary vertex
a of the complex X and consider it as an elementary zero-dimensional
chain (cycle). We shall prove that any other zero-dimensional chain is
homologous to a chain rna, where m is an integer. We shall prove that
elementary zero-dimensional chains corresponding to any two vertices
a and 13 of the complex X are homologous. Since the complex X is
linearly connected, one can always construct a continuous path "f such
that consists of one-dimensional edges of simplexes, starts at the vertex
Fig. 1.2.5 a and ends at the vertex 13 (Fig. 1.2.5). This path can be regarded as a
one-dimensional chain. To this end it suffices to prescribe orientation to
each of its links (we indicate it with arrows going from the vertex a
1.2 Simplicial Homology Groups of Simplicial Complexes (Polyhedra) 33

towards the vertex (3) and then to assign to each link, i.e. to each one-
dimensional simplex, an integer equal to unity. As a result, we obtain a
chain, = ..11 +... +..11, where N is the number of edges composing the
path. Let us find the boundary of this chain. Consider an interior vertex Interior vertex
ai of the path, i.e. a vertex which is the end of a certain edge ..11 and
the beginning of the subsequent edge ..11+1 (Fig. 1.2.5). Then we have:
8..11 = ai - ai-\, 8 ..11+1 = ai+1 - ai. Therefore,

Consequently, each interior vertex will enter the sum twice, once with the
"+" and then with the "-" sign, which entails mutual cancellation of these
two summands. This does not happen only to the two boundary vertices Boundary vertices
a = ao and (3 = aN, i.e. to the initial and terminal points of the path ,.
Hence, 8, = aN - ao =(3 - a. This means that the zero-dimensional chain
(3 - a is homologous to zero (it is represented as the boundary of some
one-dimensional chain ,). This implies our statement. Indeed, since any
zero-dimensional vertex (with coefficient 1) is homologous to the chosen
vertex a, it follows that any zero-dimensional chain I:f=1 ai..1~ = c is
homologous to the zero-dimensional chain rna = (I:f=1 ai) a. Thus, to
calculate the zero-dimensional group Ho(X) it suffices to consider only
one vertex a. From this we obtain that the group Ho(X) of the' connected
complex X is the same as that of the complex consisting of a single point Homologies of a circle
a. But the latter group is isomorphic to Z all the cycles here have the form
rna, where m is an arbitrary integer (there are no nonzero boundaries).
From the above arguments one can immediately deduce the proof
of the fact that the group Ho(X) of the disconnected complex X is iso-
morphic to 7!}, where q is the number of arcwise connected components.
We shall not do it here and leave the proof to the reader.
2) Homologies of a circle. Let us consider a simplicial partition of

,i
a circle 8 1. Clearly, it is given by the set of vertices ao, ... ,aN and the
one-dimensional simplexes confined between neighbouring vertices.
We prescribe a clockwise orientation to all these one-dimensional edges
(Fig. 1.2.6). The zero-dimensional homology group Ho(8 1) is already Fig. 1.2.6
known to us - it is isomorphic to Z. The groups Hi(8 1), where i > 1,
are equal to zero since the complex 8 1 is one-dimensional.
34 1. Polyhedra. Simplicial Complexes. Homologies

One-dimensional and It remains to calculate the group HI(SI). We state that HI(SI) = z.
two-dimensional homology An arbitrary one-dimensional chain has the form e = ao'Yo + ... +an'YN,
groups of 2-manifolds where aj are integers. In what case is such a chain a cycle? Calculating
the chain boundary, we obtain
oe = ao(al - ao) +al (a2 - al)+'" +aN(ao - aN)
We remove the brackets amd group the like terms. As a result, we are
led to the following zero-dimensional chain:
oe=(-ao+aN)ao+(ao-aJ)al + ... +(aN-I-aN)aN .
2-dimensional topological Thus, the one-dimensional chain e is a cycle if and only if oe = 0, Le. if
manifold is a polyhedron the coefficients ai satisfy the following system of equations:
-aD + aN = 0, ao - al = 0, ... , aN-I - aN =0 .
From this we obtain ao = al = ... = aN-I = aN. Thus, the chain e is
a cycle if and only if all the coefficients are equal to one another. In
other words, we habe described all one-dimensional cycles. They have
the form z = abo +... +'YN), where a is an arbitrary integer. The group
Sphere, torus, of one-dimensional boundaries BI(SI) is equal to zero since there are
sphere with handles no chains of dimension two. Therefore, the group of one-dimensional
homologies HI (Sl) coincides with the group of one-dimensional cycles
ZI (Sl). Since all one-dimensional cycles have the form a (E 'Yi), where
a E Z, it follows that the one-dimensional homology group is isomorphic
to Z. Thus, for a circle we have: Ho = Z, HI = Z, Hi = 0 for i 0, 1. r
3) Two-dimensional homology groups of two-dimensional polyhe-
dra, which are manifolds without boundary. We shall say that a two-
dimensional topological manifold has no boundary ifeach of its points
~
..::::..:...
::,. ..
:::': has a neighbourhood homeomorphic to an open two-dimensional disc (the
::
'
:'. '"
.: .. ...... .
centre of this disc is a point). Any two-dimensional topological manifold
is a polyhedron. We shall for simplicity examine only those manifolds

~CJ32
which are finite polyhedra. For example, a sphere, a torus, a sphere with
handles are manifolds without boundary. To demonstrate this defmition,
Fig. 1.2.7 Fig. 1.2.7 gives examples of manifolds with boundary. Points lying on
the boundry (boundary points) differ from interior points in that they
possess a neighbourhood homeomorphic to half a disc, the point itself
lying on the diameter of this half disc. A manifold without boundary has
no such points.
1.2 Simplicial Homology Groups of Simplicial Complexes (polyhedra) 35

So, suppose we are given a finite polyhedron which is a manifold


without boundary. Then its triangulation possesses the following impor-
tant property: each one-dimensional edge (a one-dimensional simplex) is
incident to two and only two adjacent triangles. In other words, each Adjacent triangles
edge enters the composition of exactly two triangles (glued along it)
(Fig. 1.2.8). Clearly, we essentially employ here the absence of boundary
in the manifold. If the manifold had boundary, there would exist an edge
lying on the boundary and incident only to one triangle (approaching it
from within the manifold) (Fig. 1.2.8).
We call a manifold oriented if each two-dimensional simplex of
its triangulation can be so oriented that any adjacent triangles incident
along the edge (having a common edge) induce on it opposite orientations
(Fig. 1.2.8). Otherwise a manifold is called non-oriented. It is character-
ized by the fact that for any way of orientation of triangles there exists
at least one edge on which adjacent triangles induce one and the same Fig. 1.2.8
orientation.

Theorem. The two-dimensional homology group H2(M) of a finite con- Orientable and
nected polyhedron which is an orientated manifold without boundary is non-orientable manifolds
isomorphic to a group of integers Z. If the manifold is non-oriented, the
group H2(M) is equal to zero.

Proof. Suppose the manifold is at first orientable. We consider its tri-


angulation and orient all the triangles L1l in such a way that on each
edge adjacent triangles induce opposite orientations. An arbitrary two-
dimensional chain has the form e = E~1 aiL1f, Our goal is to describe
all two-dimensional cycles. Calculation the boundary of the chain e, we
obtain ae =E aia L1f. As a result, we have obtained a one-dimensional
chain. The chain e will be a cycle if and only if the one-dimensional
chain ae is equal to zero on each one-dimensional simplex (edge of tri-
angulation). We take an arbitrary edge ,11. It is incident only to two
triangles L1l and .1; and is their common edge. By the condition, both
the triangles induce opposite orientations on the edge ,11 (Fig. 1.2.9). Opposite orientation
Consequently, the edge ,11 will enter the sum ae (after grouping like
terms) with the coefficient ai - aj, where ai and aj are the values of
the two-dimensional chain e on the triangles ,1; and L1J, respectively.
36 1. Polyhedra. Simplicial Complexes. Homologies

Since Be = 0, it follows that ai = aj. Thus, we have proved that a two-


dimensional chain e is a cycle if and only if it acquires equal values on
any two adjacent triangles. But the manifold being connected, any two
triangles t and T can be joined by a chain of adjacent triangles (Fig. 1.2.9).
The chain e will, accordingly, assume one and the same value on all the
triangles of triangulation. Hence, e = a . (~i ,1;). Thus, we have com-
pletely described all cycles in two dimensions. It turns out that the group
Z2(M) is isomorphic to the group Z seing that the coefficient a can take
on arbitrary integer values. So long as the manifold is two-dimensional,
it has no nonzero three-dimensional chains, and therefore the group of
boundaries B2(M) is equal to zero. From this it is immediate that the
homology group H2(M) is equal to Z. This proves the theorem in the
first case.
Suppose now the manifold is non-orientable. We fix orientations of
all triangles of triangulation. We again write a two-dimensional cycle e
Fig.1.2.9a
in the form ~~l aiL1f. Repeating the previous arguments we can see that
if two adjacent triangles L1r and ,1] induce distinct orientations on their
common side, then ai = aj, whereas if they induce one and the same
orientation, then ai = -aj. Consequently, by virtue of connectedness of
the manifold all the coefficients aj are equal to one another in the absolute
value, i.e. lail = lajl for any i and j. We take two adjacent triangles ,1;,
,1; and orient them so that they induce distinct orientations on their
common edge. Then ai = aj. We take successively triangles adjacent to
the two preceding ones and also orient them in a compatible way (as long
as possible). Then on all these triangles the cycle e will acquire the value
equal to the initial value ai. But since the manifold is non-orientable, the
moment will necessarily come when two adjacent triangles will meet that
induce one and the same orientation on their common edge (Fig. 1.2.9).
Given this, the cycle z assumes one and the same value ai on both
Fig.1.2.9b triangles. On the other hand, the equality ai + ai = 0 must hold. From
this we obtain that ai = O. Therefore, any two-dimensional cycle on a
non-orientable manifold is equal to zero, which completes the proof of
the theorem.

We shall give examples of orientable and non-orientable manifolds.


A plane, a square, a sphere, a torus, a ring, a cylinder, spheres with
1.2 Simplicial Homology Groups of Simplicial Complexes (Polyhedra) 37

handles are orientable manifolds. A Mobius strip, a projective plane, a


Klein bottle are non-orientable manifolds. The reader may verify it.
Thus, if a polyhedron is a two-dimensional closed manifold, we have
calculated its zero- and two-dimensional homology groups. It remais to
calculate the one-dimensional homology group. The argumentation will
be more complicated, although they are essentially elementary as well.
We shall define the concept of cycle support which will be useful for Cycle support
our further purposes. Let z be some k-dimensional cycle. The union of
all k-dimensional simplexes on which the cycle z takes nonzero values
will be called the support of the cycle.
4) Auxiliary lemma on deformation of one-dimension cycles. Lemma on deformation
of the cycles
Lemma. 1) Let in a polyhedron these exist a one-dimensional cycle z
assuming values a, b, c respectively on oriented edges t, T, (J (Fig. 1.2.10)
which are sides of a two-dimensional simplex (triangle) ,12. Then the
cycle z is homologous to another cycle z' which on the edges t, T, (J
assumes respectively the values 0, b + a, c + a. Given this, the values
of the cycles z and z' coincide on all the other edges of triangulation
(Fig. 1.2.10). Figuratively speaking, the cycle z can be taken out of the
edge t.
2) Let in a polyhedron there exist a one-dimensional cycle z as-
suming values a, b, c respectively on oriented edges t, T, (J (Fig. 1.2.10)
which are sides of a triangle ,12. Let the cycle z be equal to zero on all
the other edges of triangulation incident to the vertex A of the triangle
,12 and different from the edges T and g. Then necessarily b = c and
the cycle z is homologous to another cycle z' which on the edges t, T, (J
assumes respectively the values a + b, 0, O. Given this, the values of the
cycles z and z' coincide on all the other edges of triangulation. In other
words, the cycle z can be taken from the sum of the edges T and (J, and
its value b can be directed along the edge t.
3) Let in a polyhedron there exist a one-dimensional cycle z as-
suming a certain value a on an edge t of the polyhedron one of whose
vertices is indicent only to such edges of the polyhedron on which the
cycle z is equal to zero (Fig. 1.2.10). Then a = 0 and the cycle z is
homologous to another cycle z', in the composition of whose support the
edge t does not enter. Given this, the values of the cycles z and z' coin-
38 1. Polyhedra. Simplicial Complexes. Homologies

cide on all the other edges or triangulation. In other words, the cycle z
can be taken from the edge t.

Proof of the lemma Proof. Let the edges be oriented as shown in Fig. 1.2.10. We shall
calculate the difference z - z' of the cycles z and z'. From Fig. 1.2.11 it
seen that as a result we obtain a cycle assuming the values a on t, -a on
7, and -a on fl. We shall consider a two-dimensional chain e acquiring
the value a on a triangle Ll and equal to zero on all the other triangles
of triangulation. Let us calculate its boundary 8e. Clearly, we obtain a
Edges of triangulation one-dimensional chain equal to zero on all the edges of triangulation
different from t, 7, fl and equal respectively to a, -a and -a on these
\ l-l' / \ q ,.
latter edges. Here we make use of the fact that the oriented boundary of
the triangle Ll consists of edges t, - fl, and -7 (Fig. 1.2.11), and therefore,
\v"~/
-
()
-q
~
q 0 (8e)(t) = +a, (8e)( -fl)) = a, (8e)( -7) = -a. Thus, the cycles z and z'
are homologous since z - z' = 8e. The proof of the second and third
i d(a.t/-) =!
parts of the lemma is carried out in a similar way (see Figs. 1.2.10 and
1.2.11). We shall not dwell on this proof and only point out that the
equality b = e in item 2 the equality a = 0 in item 3 follow from the
condition 8z = O. This statement is based on the fact that the vertex A is
incident only to two edges of triangulation on which the cycle acquires
a nonzero value, namely, to the edges 7 and fl. The other edges, incident
to A, make a zero contributiop. to the zero-dimensional chain 8z. This
Fig. 1.2.11 completes the proof of the lemma.

This lemma can be imparted a clear geometrical meaning. It turns


out that we can deform the initial cycle z by displacing its support along
the surface. We replace its edge which enters into some triangle by the
sum of two other edges of this triangle, and the value a which was earlier
ascribed to the edge t ''transits'' onto these edges and is added to the for-
mer values of the cycle on these edges. As a result, the cycle z is replaced
Elimination of the edge to the one homologous to it, and the edge t can be eliminated from its
composition since on this edge the cycle z' acquires zero value. Simi-
larly, if a certain triangle has a "free" vertex, i.e. such a vertex to which
no other nonzero edges of the cycle are incident (except for two sides of
the triangle), then the cycle can be taken from the triangle through defor-
mation of the sum of two sides onto the third side (Fig. 1.2.12). In other
words, we can continuously deform one-dimensional cycles about the
1.2 Simplicial Homology Groups of Simplicial Complexes (polyhedra) 39

polyhedron sliding along two-dimensional simplexes (without changing


the homology class of the cycle). Figure 1.2.12 demonstrates the pro-
cess of cycle extrusion along the triangle and its replacement by the sum
of two other sides of the triangle. We shall apply this simple but use-
ful lemma to calculate one-dimensional homologies of two-dimensional
manifolds. It is useful to think of a one-dimensional cycle as a closed
rubber thread located on the manifold. We deform the cycle by expand-
ing or contracting this thread without breaks (glueings are admitted). We
can thus essentially simplify the cycle by disposing it on the surface in
an optimal way.
All these operations can be visually interpreted in terms of the sup-
port of the one-dimensional cycle z. According to the lemma, we can
subject the support to the following operations (surgery).
1) Any edge of the support entering in the composition of a two-
dimensional simplex can be replaced by the sum of two other edges of
this simplex (Fig. 1.2.12) with a corresponding change in the numerical
values of the cycle.
2) Some pairs of edges converging at a free vertex of the triangle
can be replaced by one edge (Fig. 1.2.12).
3) An edge of the support protruding sideways and having a free
end can be removed (Fig. 1.2.12).
These operations can be called combinatorial disassembly (simplifi-
cation) of the cycle. Clearly, each of the operations 1-3 decreases the Fig. 1.2.12
number of edges in the cycle support and therefore we actually simplify
the support.
If we digress from triangulation and consider the support as a one-
dimensional set consisting of segments of polygonal lines on a manifold,
then the operation of support simplification introduced above can be
interpreted in another manner. If several polygonal lines in some region
are rather close and parallel to one another (but, perhaps, differently
oriented), this system can be replaced by another one (Fig. 1.2.12) by
glueing together all the previous polygonal lines. In doing so, one should Combinatorial disassembly
sum up (with allowance made for signs) all the numbers which stood on of the cycles
the previous polygonal lines and ascribe the result obtained to the new
segment. One can simplify the cycle by cutting from it (or pulling onto it)
some "branches", i.e. polygonal lines with free ends which come from a
40 1. Polyhedra. Simplicial Complexes. Homologies

Polygonal lines certain point of the cycle and do not return onto this cycle. Thus, a system
with free ends of rubber threads representing a cycle can be continuously deformed
about the manifold, by glueing the threads and simplifying the cycle
support. For example, a closed curve drawn on a torus and going round
it twice along the parallel (and evidently representing a one-dimensional
A#\~ cycle) can be continuously deformed into a curve which is also a parallel

V@]~O}tJ
11; .;
but taken with coefficient 2 (Le. into a parallel passed twice).
5) Homologies of a sphere. We shall consider an arbitrary fmite
,.,w ......
triangulation of a two-dimensional sphere (Fig. 1.2.13). Let z be an ar-
Fig. 1.2.13 bitrary one-dimensional cycle. We state that it is homologous to a zero
cycle, i.e. to a cycle which is identically zero on all edges of triangula-
tion. We apply the preceding lemma. It states that one can always remove
any edge ("erase" it) from the composition of the cycle z by adding the
numerical value given initially on this edge to two adjacent edges of the
triangle which contained this edge. As a result, one obtains a new cycle
homologous to the initial one. We start this operation with an arbitrary
Homologies of a 2-sphere edge. We erase it, redistribute the number ascribed to it between the ad-
jacent edges, take the next edge, erase it, etc. Repeating the process a
fimte number of times and combining it with operations 2 and 3 (see the
lemma), we run over all the edges of triangulation (they are finite). In
Case of the penultimate step we obtain a cycle equal to zero everywhere except
n-dimensional sphere on the edges of one triangle, on which the cycle assumes vales a, b, c.
But since we are dealing with a cycle, a = b = c. Consequently, in the
last step, extruding one of the edges onto the sum of two other edges. we
liquidate all the three numbers and as a result obtain a cycle homologous
Deformation of to the initial one but equal to zero on all the edges. We have proved that
multidimensional cycles the one-dimensional homology groups of a sphere is trivial (equal to
zero). Thus, HO(S2) = 'l, HI (S2) = 0, H2(S2) = 'l. We invite the reader
to prove that the homology groups of an n-dimensional sphere have the
form: Ho(sn) = 'l, Hi(sn) = 0 for 1 ~ i ~ n - 1, Hn(sn) = 'l. At the
same time the reader will have to prove a multidimensional analogue of
the lemma on cycle deformation. From the proof of our lemma one can
see that one can similarly deform supports of multidimensional cycles by
extruding them inside a multidimensional cycle and replacing one face
of the simplex by the sum of the rest of its faces (with altered numerical
values).
1.2 Simplicial Homology Groups of Simplicial Complexes (Polyhedra) 41

6) Homology of a torus. We represent a torus in the form of a


square with opposite sides identified (Fig. 1.2.14). Consider triangulation
of a torus depicted in Fig. 1.2.14. We have deliberately indicated orienta-
tions of all the simplexes to show that a torus is an orientable manifold.
One can see that all the orientations are coordinated, i.e. adjacent tri-
angles induce different orientations on their common sides. Suppose we
are given a one-dimensional cycle z on a torus. We shall prove that it
is homologous to a cycle of the form ma + n(3, where a and (3 are the
two cycles shown in Fig. 1.2.14 and m and n are some integers. More
precisely, the cycle a is equal to a+b, where a and b are oriented edges Ai e
(Fig. 1.2.14), and the cycle (3 is equal to c + d. We shall apply the three Fig. 1.2.14
above-described operations of cycle simplification (deformation).
As a result (Fig. 1.2.14) we can take the support of the cycle z from
all the interior edges of triangulation which come from the centre of the
square. We obtain a cycle z' whose support consists only of boundary
edges of the square. One cai:L see that a = a +b and (3 = c +d. It remains
to prove that the cycle z' = m(a +b) +n(c +d) is homologous to zero if
and only if the coefficients m and n are equal to zero. Indeed, suppose
that at least one of the numbers m and n is nonzero and that there exists Homology of a torus
a two-dimensional chain q such that 8q =z'. Since the cycle z' is equal
to zero on all interior edges of triangulations (incident to the centre of the
square), it follows that the chain q takes necessarily equal values on all
adjacent (by a common edge) triangles. Consequently, the chain q takes
one and the same value A on all the triangles of triangulation. The point Values of the chain
is that any pair of triangles is joined by a chain of triangles adjacent by on the edges of
a common edge. But in such a case the boundary aq of the chain q takes triangulation
zero on all the edges a, b, c, d. Indeed, e.g. two copies of the side a (on
opposide sides of the square) evidently induce reverse orientations of
the square, and therefore when taken with coefficients equal in absolute
value but different in sign, they annihilate each other in the boundary of
the chain q. These arguments are also applicable for the edges b, c, d. As
a result we obtain 8q = O. But since 8q = z', we have z' = O. Therefore,
m(a + b) +n(c +d) =O. It remains to make sure that the cycles a =a +b
and (3 = c +d are linearly independent. But this is obvious because the
cycle (3 is equal to zero on edges c, d, and the cycle (3 is equal to zero
on edges a, b. From this we obtain that m = n = O. One should bear in
42 1. Polyhedra. Simplicial Complexes. Homologies

mind that all the four vertices of the square, AI,A2,A3,~, represent
actually one and the same polyhedron vertex. This fact is dictated by the
rule of identification of the sides of a square in the case of a torus. This
proves the statement.
Thus, each one-dimensional cycle is homologous to a cycle of the
form ma+n{3, the cycles a and {3 being linearly independent. Therefore,
the one-dimensional homology group HI (T2) of the torus is isomorphic
to the direct sum Z EB Z two copies of the group Z.
Returning to the model of a torus in the form of a ''roll'' (Fig. 1.2.14)
we can see that the generator a corresponds to the torus parallel and the
generator (3 to the torus meridian. As has been proved above, any other
cycles on the torus are a linear combinations ofthe parallel and meridian
taken with arbitrary integer coefficients.
7) Homologies of the sphere with 9 handles. It is now clear how
one should calculate the one-dimensional homology group of a sphere
with an arbitrary number of handles. Recall that a torus is a sphere with
a single handle. A sphere with 9 handles is convenient to represent as
Fig. 1.2.15 shown in Fig. 1.2.15. One can see that the consideration is largely reduced
to a multiple repetition of the construction which we have realized on a
torus. To this end we construct a convenient triangulation of a sphere with
9 handles. We shall cut the surface curvilinearly, as shown in Fig. 1.2.15.
As a result, we obtain a flat polygon with 4g sides. On a handle with
number i we denote the parallel by ai and the meridian by {3i. Then the
sides of the polygon obtained are numbered as follows (we write down
the letters in the order of their appearance in a clockwise detour on
the boundary: 0'.1 {31 all {31 1... a g{3g a; I{3; I. We triangulate the polygon
(and thus the surface) as shown in Fig. 1.2.16. Clearly, we can now repeat
the arguments used above for the torus. First we prove that each cycle z
is homologous to a cycle z' which has the form

z' mlal +nl{31 + .. ·+mgag+ng{3g


ml(al +bl)+nl(ci +dl)+ .. ·+mg(ag+bg)+ng(cg+dg)

Here mj, ni are integers. Next, we prove, as in the case of the torus,
that all the cycles aj = aj + bi and (3j = Cj + dj, where 1 ~ i, j ~
g, are linearly independent. We therefore point out in an explicit form
Fig. 1.2.16 the basis in the group H1(Mg), where Mg is a sphere with g handles.
1.2 Simplicial Homology Groups of Simplicial Complexes (polyhedra) 43

Namely, the basis consists of the cycles ai, ... , a g and (31, ... , (3g. Thus, Homologies of a
HI(Mg) = Z E!1'" E!1 Z (29 times), i.e. HI (Mg) = Z29. Summarizing the sphere with handles
results obtained above, we now can formulate the following theorem.

Theorem. Integer-valued homology groups of a sphere with 9 han-


dles, where 9 is a non-negative integer, have the form: Ho = Z, HI =
Z2g ,H2 = Z.

8) Homologies of a projective plane Rp2. Let us consider the tri-


angulation of a projective plane shown in Fig. 1.2.17. It is distinct from
the triangulation of a torus. Although in both cases we triangulate a
square, glueings on its sides are organized differently. Given this, we
have prescribed specially the orientation on two-dimension~simplexes,
as on the triangluation of a torus, in order that non-orientability of the
projective plane might be clearly seen. Indeed, on the square bound-
ary, on the sides a, b, c, d, identical orientations are induced. This fact
is indicated in Fig. 1.2.17 by the "+" signs. Proceeding by analogy with
the torus we obtain that anyone-dimensional cycle z is homologous to
the cycle z', whose support consists of boundary edges of the square. Fig. 1.2.17
In other words, z rv ma + n(3 = m(a + b) + n(c + li) = z', where
m and n are integers. Here a and (3 (separately) are not cycles so
long as oa = A2 - Al f 0, 0(3 = A3 - A2 f O. In the case of torus a
and (3 are on the contrary cycles in so far as on the torus we have
Al = A2 = A3 = AJ. In the case of projective plane there is, however, an
additional circumstance which makes our further calculations different
from those in the case of torus. Indeed, let us calculate the boundary of Homologies of a
the cycle z'. On the one hand, this is a zero-dimensional chain of the form projective plane
oz' = o(ma+n(3) = o(ma+mb+nc+nli) = m(A2 - AI)+n(A3 - A2). But
by the rule of identifications of sides of a square only two of the vertices
A I,A2,A3,AJ (Fig. 1.2.17) are distinct since Al = A3 and A2 = A4•
Consequently, 0 = oz' = (m - n)A2 + (n - m)AI. Considering that Non-orientability
Al f A2, we have m = n. Thus, we have proved that on a projective of projective plane
plane anyone-dimensional cycle is homologous to a cycle z' of the form
(m + n)(a + (3) = q(a + (3) = n, where "1 = a + (3 and q is an arbitrary
integer. Next, we state that the cycle "1 taken twice is homologous to
zero, i.e. 2"1 rv O. Indeed, let us consider a two-dimensional chain T
which takes the value 1 on each oriented triangle of our triangulation.
44 1. Polyhedra. Simplicial Complexes. Homologies

Calculating its boundary, we obviously obtain from Fig. 1.2.17 the rela-
Projective plane tion 8T = 2a + 2b + 2c + 2d = 2"1, and the statement follows. We have
and Klein bottle proved therefore that the group HI (RP2) is isomorphic to the group Z2
whose generator is the cycle "I depicted in Fig. 1.2.17, i.e. "I = a+b+c+d.

Theorem. Integer-valued homology groups of a projective plane have


the form Ho = Z, HI = Z2, H2 = O.

9) Homologies of a Klein bottle. We consider triangulation of a


Klein bottle K illustrated in Fig. 1.2.18. One can see that this manifold
is non-orientable. As in the preceding cases, we immediately obtain that
anyone-dimensional cycle z is homologous to a cycle z' of the form
z' = rna: + n(3, where rn and n are some integers and a: and (3 are
cycles, and a: = a + b, (3 = c + d. We have: 8a: = 8(3 = 0 since all the
four vertices of the square represent actually one vertex on the Klein
bottle, i.e. Al = A2 = A3 = A4 (as on a torus). We shall examine a
two-dimensional chain T which takes the value 1 on all the oriented
Fig. 1.2.18 triangles (fig. 1.2.18). Calculating its boundary, we obtain 8T =a: +(3 -
a + (3 = 2(3. Accordingly, a double cycle is homologous to zero. Thus,
the one-dimensional homology group is isomorphic to the direct sum of
two Abelian groups Z E9 Z2, where the subgroup Z is generated by the
generator a (an infinite-order cycle), while the subgroup Z2 is generated
by the generator (3 (a second-order cycle, i.e. 2(3 rv 0).

Theorem. Integer-valued homology groups of the Klein bottle have the


form Ho = Z, HI = Z E9 Z2, H2 = O.

10) Homologies of non-orientable two-dimensional closed surfaces.

We shall consider the polygon depicted in Fig. 1.2.19. It is a na-


tural generalization of the polygon which determines a projective plane
(Fig. 1.2.17). The reader may verify that identifying edges on the poly-
gon boundary in accordance with the indications (letters and arrows),
we shall obtain a two-dimensional closed compact manifold. Besides,
this manifold will be non-orientable. Omitting the calculational de-
Fig. 1.2.19 tails which we have repeatedly demonstrated already, we obtain that
anyone-dimensional cycle z is homologous to a cycle z' composed of
1.2 Simplicial Homology Groups of Simplicial Complexes (polyhedra) 45

edges - cycles 0:1,"" O:k taken with arbitrary integer coefficients, i.e.
z' = ml 0:1 +... +mkO:k. Given this, the cycle 2(0:1 +... +O:k) is homo-
logous to zero since it is the boundary of a two-dimensional chain which Homologies of
takes the value 1 on each oriented triangle. Consequently, for the gener- non-orientable 2-surfaces
ators in a one-dimensional homology group one can take the following
cycles: 0:1, ... , O:k-I, h, where h = 0:1 + ... + O:k-I +O:k. Then 2h 0,tv

and the cycles 0:1, ... , O:k-I are free. As a result, the homology group
HI is isomorphic to the direct sum Zk-I Ell Z2.

Theorem. Integer-valued homology groups of a non-orientable two- Homologies with coefficients


dimensional closed surface have the form: Ho = Z, HI = Zk-I Ell Z2, in the group 2:2
H2=0.

We have hitherto calculated integer-valued homologies of polyhedra.


As a useful (and simple) exercise we advise that the reader calculate
homologies of all the above-mentioned polyhedra with coefficients in
the group Z2 or more generally in the group Zp for an arbitrary prime
p. We shall only present the result withOl~t going into detail Of the proof
(although we have in fact presented all the necessary material).
Homology groups of a two-dimensional manifold M with coeffi-
cients in the group Z2 have the following form.
1) In the case of a sphere with 9 handles we have: Ho = Z2, HI = Sphere with handles
Z2 Ell ... Ell Z2(2g times) = z~q, H2 = Z2.
2) In the case of a non-orientable manifold we have: Ho = Z2, HI =
Z2 Ell··· Ell Z2(k times) = Z~, H2 = Z2.
We should emphasize the following: while the two-dimensional
integer-valued homology group of a non-orientable manifold is equal
to zero, the two-dimensional homology group of the same manifold, but Sphere with Mobius strips
with coefficients in the group Z2, is non-zero and equal to Z2.
46 1. Polyhedra. Simplicial Complexes. Homologies

1.2.7 Visual Material

Figure 1.2.20 represents an infinite polyhedron composed of three-


dimensional simplexes (tetrahedra). A similar picture is obtained by un-
folding a ball triangulated so that the simplexes have bases on the ball
boundary (on the sphere) and one vertex at the ball centre. In Fig. 1.2.21
Algebraic chain one can see an algebraic chain composed of discs = cells endowed with
numerical coefficients (numbers are conditionally represented by various
figures). The condition that the chain is a cycle implies that all these
numbers are equal, i.e. all the figures must be identical. In case the
figure~ are distinct (as in the Fig. 1.2.21), the chain is not a cycle.

Page 47, Fig. 1.2.20. Simplifical subdivision of a topological space; spines


of 3-manifolds. Valley of ancient kings; egyptian myth on the forgotten
tombs of gods and pharaohs
Page 48, Fig. 1.2.21. Random walk and random sequences. Homology
chains with coefficients in Abelian group. Medieval legend on the Flying
Hollander; Scandinavian myths
1.3 General Properties of Simplicial Homology Groups 49

1.3 General Properties of Simplicial Homology Groups.


Some Methods for Calculating Homology Groups

1.3.1 Incidence Matrices Incidence matrices

We shall consider an arbitrary polyhedron X and its arbitrary trian-


gulation. We shall number all of its oriented simplexes by symbols
Llf, where k indicates the simplex dimension and i shows its ordi-
nal number (in a particular dimension), i.e. 1 ::; i ::; G:k. Here G:k
is the number of k-dimensional simplexes in the polyhedron. Besides,
1 ::; k ::; n, n = dimX, i.e. n is the polyhedron dimension. We assume
for simplicity that the polyhedron is finite, then G:k < 00. From what
follows it will be clear how we extend most of the constructions to the
case of infinite polyhedra. Let us consider the boundary operator 8. For
each simplex LlJ+l we obtain the relation 8LlJ+l = L~kl atLlf, where
afj are arbitrary integers. Given this, we have 1 ::; j ::; G:k+l, 1 ::; i ::; G:k.
The number afj is equal either to +I or to -1 depending on the orien-
tation with which the simplex Llf enters in the boundary of the simplex
LlJ+l. In other words, afj is equal to +1 if the orientation given on LlJ
coincides with that induced on it by the simplex LlJ+l, and otherwise at
is equal to -1. The number at
is called the incidence coefficient of the Incidence coefficients
simplex Llf and simples Llj+l. In case the simplexes Llf and LlJ+l are
not incident (i.e. if the simplex Llf is not a face of the simplex LlJ+ 1), Boundary operator
then by definition afj = O. Clearly, the numbers afj are organized in
a natural way into a rectangular matrix which we denote by Ek. It is Rectangular
convenient to write this matrix as incidence matrix
50 1. Polyhedra. Simplicial Complexes. Homologies

k-dimensional This matrix has Cik rows and Cik+I columns. It is called a k-dimensional
incidence matrix incidence matrix. Thus, with each triangulation of the polyhedron there is
associated a set of incidence matrices EO, ... , En-I, where n = dim X.
Let us consider the simplest examples.
In Fig. 1.3.1 one can see two polyhedra: a segment and a triangle. In
the case of a segment we have a single incidence matrix EO of the form

.1 1
EO=.10 -1
1
.1~ +1

For a triangle we have two incidence matrixes: EO and EI, namely

Ef< .1 11 .1i .131 .12

Fig. 1.3.1 EO= .101 0 -1 -1 EI= .1 11 +1


.10
2 -1 0 +1 .121 -1
.10
3 +1 +1 0 .131 +1

As an exercise the reader may compose three incidence matrices cor-


responding to the standard triangulation of a three-dimensional simplex
Tetrahedron (tetrahedron).

1.3.2 The Method of Calculation of Homology Groups Using


Incidence Matrices

What is the role of incidence matrices? It turns out that a simplicial com-
Polyhedron triangulation plex (co"esponding to a given polyhedron triangulation) is completely
(and uniquely) determined by the incidence matrices.
Incidence matrices Indeed, knowing these matrices and taking an arbitrary simplex .1j
of a simplicial complex we uniquely restore to which simplexes of dimension n - 1 it is inci-
dent. Continuing this process in the order of decreasing dimension, we
ultimately restore the whole scheme of simplicial complex, i.e. find out
which of its simplexes are incident and which are not. Consequently,
all the properties of a simplicial complex are laid in its incidence ma-
trices and must be calculable if these matrices are known (given). This
also refers to the simplicial homology groups of a complex. To put it
1.3 General Properties of Simplicial Homology Groups 51

differently, knowing the incidence matrices of a simplicial complex (cor-


responding to a given polyhedron triangulation), we can calculate all
its homology groups. We shall show how this can be done. This will
just give us one of the ways of algorithmic calculation of simplicial Computer program
homology groups. In other words, this technique makes it possible to
write a computer program and then, applying concrete incidence matri-
ces of polyhedra, calculate algorithmically all the homology groups of System of equations
the polyhedron. for incidence matrices
As we already know, a boundary operator squared is zero, i.e. & = o.
From this we obtain the following system of equations for incidence
matrices of neighbouring dimensions:
=~~kI a~.aLl~ = ~~I . a~· . ~akI-l . Llk-I
LJt- tJ t LJ,- tJ LJq= q Algorithmical calculation
of the homological groups
=~ak-l . (~?'k. ~.. ~-I). Ak-I = 0
LJq= I LJt= I atJ a,q Llq .

Since the simplexes Ll~-I are linearly independent (as elementary chains),
equality to zero of their linear combination with some coefficients im-
plies that all these coefficients are equal to zero. Therefore, we are led
to the system of equations: Ef:i afja~q- I = 0, where 1 ::; j ::; ll!k+ I, 1 ::;
q ::; ll!k- I. Clearly, for a fixes k this system is equivalent to a single
matrix equation, namely, E k- I Ek = O. Here the index k runs over the
range from 0 to n - 1. Figure 1.3.1 shows that muliplying the ll!k- I X ll!k
matrix Ek-I by the ll!k x ll!k+I matric Ek, we obtain a new ll!k-I x ll!k+I
matrix composed entirely of zeros.
The system of matrix equations obtained writes, in the incidence
matrix language, the fact that the square of the boundary operator a Simplification
is equal to zero. Suppose that instead of the elementary chains chosen of incidence matrices
above we have passed over to another basis in the chain space, using
a nondegenerate linear transformation. Then the incidence matrices Ek
will, of course, transform to become some new matrices Ek. But it is
readily seen that the matrix equation E k - I Ek = 0 remains unchanged,
i.e. the identity Ek- I Ek = 0 holds as before. Indeed, passing over to
another basis in the chain space does not alter the geometric fact that the
square of the boundary operator is equal to zero.
This simple observation offers the opportunity to simplify incidence Change the basis
matrices by changing the basis in the chain spaces. We wish to reduce in the chain groups
these matrices to the simplest, canonical form in order that we could
52 1. Polyhedra. Simplicial Complexes. Homologies

calculate quickly homology groups. Our transformations will consist of


successive steps, and in each step we shall change only one element
of the basis (in the chain group) in a particular dimension. We call
Elementary transformations these simplest transformations elementary (which bears analogy with the
algebraic language), and they consist in the following.
1) A simplex Llf can be replaced by a simplicial chain Llf + Ll~,
where irt.
2) A simplex Llf can be replaced by a chain -Llf.
Here we regard each simplex as an elementary chain. What happens with
incidence matrices under elementary transformations?

Different types Lemma. 1) Under type 1 transformations, the matrix tk-1is obtained
of transformations from the matrix Ek-l by adding to the i-th column of the matrix Ek-l
its t-th column. The matrix tk is obtained from the matrix Ek by sub-
straction of the i-th row from its t-th row.
2) Under type 2 transformations, the i-th column of the matrix Ek-l
and the i-th row of the matrix Ek are multiplied by -1 (i.e. all the
elements of this column and this row reverse signs).
3) On renumbering simplexes (i --t j and j --t i) the corresponding
rows (columns) exchange places.

The proof of this lemma is trivial, it reduces to a simple calculation,


and we leave it to the reader as a useful exercise which rests on the
standard material of linear algebra.
From this it follows that using elementary transformations in chain
space, one can realize elementary transformations of incidence matrices,
I,lamely, add rows, add columns, multiply a row by -1 and a column by
-1, transpose rows (columns). These operations are familiar from linear
algebra. Next, we refer to the well-known algebraic theorem stating that
Integer-valued matrices any integer-valued matrix can, using the above-mentioned elementary
transformations, be reduced to a canonical form. This canonical form is as
Canonical form follows: on the principal diagonal of the matrix (corning from the left top
comer of the matrix) there stand nonzero invariant factors whose number
is equal to the matrix rank. All the other matrix elements are equal to
zero. Recall that the diagonal elements (invariant factors) obtained are
so ordered that each of them is a divisor of a subsequent one. Invariant
Invariant factors factors are uniquely determined by the matrix and can be calculated in
1.3 General Properties of Simplicial Homology Groups 53

terms of its minors. Since we do not need here their explicit form, we
omit these formulas and only note that the very fact of their existence Algorithmic calculation
permits algorithmic calculation of all the invariant factors. of invariant factors
We now change a little the canonical form and normalize the inci-
dence matrices. It should be understood that it is insignificant for us that
the matrices EO, ... ,En-I are incidence matrices of a certain simplicial
complex. It is only essential that they satisfy the equations Ek- IEk = 0.
In other respects these matrices are arbitrary.
We begin for simplicity with the matrix EO. We reduce it to the
canonical form and then, applying elementary transformations, permute
invariant factors of the matrix EO so that they single out the right top
,0
comer of the matrix EO (Fig.1.3.2). Let be the rank of the matrix
EO. The matrices EO and EI being related by the equation EO EI =
0, the transformation of the matrix EO to the form HO automatically Fig. 1.3.2
entails the transformation of the matrix EI. Since EO EI = 0, the last
,0 rows of the transformed matrix EI must consist entirely of zeros.
Otherwise, multiplying be nonzero invariant factors, we would obtain
nonzero elements.
Now we start transforming the upper al -,0 rows of the matrix EI.
These transformations induce transformations of the first al-,o columns
of the matrix EO, and since these columns already consist entirely of
zeros, the corresponding incidence coefficients a~j will not change. It
is only one-dimensional basis chains that can change. Given this, the
matrix EO assumes its final normal form which we denote by HO. Next,
making transformations of the columns of the matrix EI (which do not
already change the matrix HO), we reduce the matrix EI to the normal
form HI in which the invariant factors occupy the same position as in
the matrix HO (they out the right top comer of the matrix). This process
can be extended to higher dimensions. As a result, we reduce all the
incidence matrices to the normal form. We shall describe in more detail Normal form
the indicated normal form vf the matrices. of the incidence matrix
The incidence matrix Ek reduces to the normal form Hk is such
that all its elements are equal to zero except for those positioned on the
diagonal segment which singles out the right top comer of the matrix
(Fig. 1.3.3). On this diagonal there first appear gk numbers cf, ... ,c~k' Diagonal of
different from unity, each of which is a divisor of a preceding one. The incidence matrix
rest of the numbers on the diagonal are equal to unity. Their amount is
54 I. Polyhedra. Simplicial Complexes. Homologies

equal to Ik - (lk. It turns out that this form of incidence matrices permits
immediate calculation of integer-valued homology groups. Let us repre-
sent a homology group Hk(X) as the direct sum of a free Abelian group
F and a fmite group T. The group F = 'lf3k is uniquely characterized
Rank of homology group by its rank {3k (by the number of free generators, i.e. infinite~order gen-
erators). The group T can be represented as the direct sum of its finite
subgroups 'lc~ E9 ... E9 'lc~k of orders cf, ...
,C~k respectively, and (as is
cf
known from algebra) the orders can always be assumed to be a divisor
Torsion coefficients of a preceding one. The numbers cf, ...
'~k are sometimes referred to
of homology group as k-dimensional torsion coefficients. Thus, the k-dimensional homology
group Hk(X) is uniquely defined by its rank and its torsion coefficients.
All these umbers appear to be calculable if one knows normalized inci-
dence matrices. There holds the following important theorem.

Theorem. Let Ilk be the number of k-dimensional simplexes of a sim-


plicial complex X (which corresponds to a certain triangulation of the
polyhedron X) and let Ik be the rank of the incidence matrix Ek. Then
the k-dimensional Betti number (3k, i.e. the rank of the integer-valued
homology group Hk(X), is equal to Ilk - Ik - Ik-I. Given this, we
assume that I-I = In = O. Next, k-dimensional torsion coefficients of
the group Hk(X) are those invariant factors of the incidence matrix Ek
Scheme of the proof which are not equal to unity. For k = n these torsion coefficients are
absent, i.e. the group Hn(X), where n = dim X, is always free (does
not have finite-order elements).

In terms of the above notation we have the identities: cf = cf, ... ,c~k
= ~k' where {lk = '{Jk;{3k = Ilk -,k -'k-I· We shall present the scheme
of the proof of this theorem. Consider k-dimensional chains of the com-
plex X and divide their basis into three groups. Chains of each group
Normalized we denote, respectively, by Ak, Bk and Ck. Any k-dimensional chain
incidence matrices is representable in the form of their linear combination. We choose the
indicated three groups according to the form of the matrices H k- I and
Hk (Fig. 1.3.3). The number of chains of the form Ck is equal to Ik-I.
Incidence equations All of them have a nonzero boundary, i.e. they are not cycles. Indeed,
the boundary of each chain of the form Ck is obtained by multiplication
of a chain of the form Ak-I by a nonzero invariant factor of the matrix
H k- I (Fig. 1.3.3). We shall write the three groups of chains introduced
above in an explicit form and indicate their number.
1.3 General Properties of Simplicial Homology Groups 55

We have:

Ak .. Ak1, .•. Ak.


""'Yk '
B k •. Bk1, ... , BPk'
k . Ck •. C1,
k •.. , C'Yk-I
k

This notation is introduced for reader's convenience in Fig. 1.3.3 on the Finite-order elements
matrices H k - 1 and Hk.
We now calculate the boundaries of chains of the form Ak and Bk.
From the form of the matrix H k- 1 it is clear that all these chains are k-
dimensional cycles since all the matrix columns corresponding to chains Generators of
of the form Ak and Bk consist entirely of zeros. Thus, oAf = 0, oBj = O. homology groups
Next, from the normalized incidence matrices one derives the following
system of equation: cf Af oCr
= 1, where cf
are nonzero numbers that
stand on the bold-faced diagonal of the matrix Hk in Fig. 1.3.3 (i.e. in-
variant factors). Consequently, all cycles of the form Afare finite-order
elements, and this order is equal to the number cf,
i.e. to the corre-
sponding invariant factor. Thus, the cycles A~, ... , ~k are generators of
fmite subgroups in the homology group Hk(X), and the order of these
Af
generators (generated by elements of cyclic subgroups) is equal to cf.
We now consider the remaining cycles of the form Bk which are
obviously equal in number to Ci.k -'Yk -'Yk-l (Fig. 1.3.3). All these cycles I ,
are remarkable for the fact that none of their linear combination with 1c I I
A I I
nonzero coefficients is a cycle homologous to zero. This is clearly seen : I
, I
from the matrix Hk, where the horizontal line composed of rows of the
form Bk consists entirely of zeros. This just means that the boundaries
of all (k +1)-dimensional chains (with our normalization of the incidence
matrix) are "outside" the chains Bk and their linear combinations. All "'\IH>'- ~

the cycles Bf are, therefore, free generators in the homology group. This fle+! P1:+1 J'f< 1-/*
completes the proof of the theorem. Fig. 1.3.3
Let us sum up the results. The basis in a k-dimensional chain group
can be divided into three parts, i.e. into chains of the form Ak, Bk
and Ck. Chains of the form Ck are not cycles, and therefore are of
no interest for us from the point of view of the calculation of the k-
dimensional homology group Hk(X). Chains of the form Ak and Bk
are cycles. Cycles Bf are free generators and cycles Aj are finite-order
generators, this order being equal to the corresponding invariant factor
of the incidence matrix.
At the same time one should realize very clearly that the practical
calculation of homologies using incidence matrices is rather cumbersome
56 1. Polyhedra. Simplicial Complexes. Homologies

since triangulatiop of even fairly simple surfaces (a torus, etc.) contains


quite many simplexes. Reduction of large integer-valued matrices to the
canonical (and normal) form is hard work, although theoretically it is
always possible, in particular, using computer. An urgent task is therefore
to speed up the obtaining of the result. Such a simplification 'is achieved
in full measure by employing the language of cell homologies which
we cannot consider in detail in this book. For a thorough presentation
see, for example, [27], [28], [24]. It turns out, however, that "traces" of
Cell homologies cell homologies are well seen and can be explained inside simplicial
homologies. We shall demonstrate it in the subsection to follow. See
also [29], [30], [31], [32].

1.3.3 "Traces" of Cell Homologies Inside Simplicial Ones

All calculational difficulties arising in the attempt of real calculation


of simplicial homology groups root in their defmition, namely, in the
Cell homologies fact that chains are the sum of simplexes taken with some coefficients.
"inside" simplicial ones Although a simplex is a simple object, there are very many of them, and
therefore it is difficult to work with such chains. But one may attempt
(and this attempt proves to be crowned with success) to unite simplexes in
Cells larger objects (blocks) which will be called celis, and the corresponding
chains will be called cell simplicial chains. Then we shall work with
Cell simplicial chains these large blocks as if they were individual "simplexes", i.e. we shall
use the old rules. This appreciably shortens the process of calculation
of homologies. It is precisely on this way that homology groups can be
defined (in terms of cell chains). We shall not do this in detail and refer
the interested reader to Refs. [27], [28], [33].
So, we choose (inside a complex X in the dimension k a finite num-
ber of simplicial complexes Qf, ... ,Q~k which are linearly independent,
i.e. the equality tJ Qf + ... + tJ,lk Q~k = 0 implies that all the numberi-
cal coefficients ti are equal to zero. This condition will be fulfilled if,
for example, no two of these complexes have common k-dimensional
simplexes. A linear combination of such complexes will be called a cell
simplicial chain. Clearly, we can speak of the boundary of a cell simpli-
Cell simpliCial cycles cial chain, ofcycles, etc. We assume the boundary of a (k+ 1)-dimensional
cell simplicial chain to be a k-dimensional cell simplicial chain. To this
1.3 General Properties of Simplicial Homology Groups 57

end it suffices to require that the boundary of each simplicial complex


of the form Q~+l be a linear combination of simplicial complexes of
the form Qj. Next, we require that for each k-dimensional simplicial
cycle (in the above-mentioned sense) there exist a cell simplicial cycle Cell complex
homologous to it. Finally, we should require that if a k-dimensional cell
simplicial cycle is homologous to zero (in the usual sense), i.e. if it is the
boundary of a certain (k + 1)-dimensional simplicial complex, then it is
also the boundary of a certain (k +1)-dimensional cell simplicial chain.
It is readily seen that in each simplicial complex one can always find Algorithm for normalizing
a set of complexes of the form Q~ satisfying all the above-mentioned of incidence matrices
conditions. Now one can start calculating homology groups using not
the former small simplexes, but larger pieces-cell simplicial chains, or
blocks. Calculating, one should use the previous recipes. The only dif-
ference is that the starting point of calculations will be not the previous
incidence matrices of an initial simplicial complex, but the new cell sim-
plicial incidence matrices defined by the boundary conditions for cell
simplicial chains. In other respects the algorithm of reducing incidence
matrices to the canonical form remains the same. We shall not here go Canonical form
deep into the theoretical detail of this procedure and invite the reader to
think over this geometrically perfectly clear and visual scheme. Rather
than abstract formal calculations we prefer the demonstration of the idea
on simple examples.
For brevity we shall refer to cell simplicial chains simply as cell Cell chains
chains and to basis complexes as cells. So, cell chains are linear combi-
nations of cells (with numerical coefficients).

We shall consider a torus depicted in Fig. 1.3.4 as a square with Cell homology groups
identified opposite sides. For the basis of cell chains we can take, for
example, the following objects:

1) One zero-dimensional cell, i.e. a zero-dimensional basis cell


chain QO, which is the only vertex. In Fig. 1.3.4 it is represented by the
four vertices of the square. Clearly, all of them are identified into one
point on the surface due to the glueings on the sides of the square.
2) Two one-dimensional cells, i.e. two one-dimensional basis cell
chains a and (3.
3) One two-dimensional cell, i.e. the two-dimensional basis cell Cell boundaries
chain Q2 is the square itself.
58 1. Polyhedra. Simplicial Complexes. Homologies

Let us calculate the cell homology groups. Clearly, the zero-dimensional


group Ho is isomorphic to Z since we have a single zero-dimensional
cell QO (which is at the same time the only zero-dimensional cycle).
Homology groups Next, anyone-dimensional cell cycle z is a linear combination of two
of a torus cell cycles O! and {3, i.e. z =mO! +n{3. We shall fmd the one-dimensional
cycles which are boundaries. We have a single two-dimensional cell,
the square Q2, and therefore an arbitrary two-dimensional cell chain is
written as pQ2, where p is an ~teger. We have: aQ2 = O! +{3 - O! -
(3 = 0, i.e. a(pQ2) = 0 for any p. Consequently, we have only a single
one-dimensional cycle which is the boundary and this is the zero cycle.
Thus, the one-dimensional homology group is isomorphic to Z EB Z,
the generators being the cell cycles O! and {3. At the same time we have
already found the two-dimensional homology group in as much as all the
two-dimensional cell chains in our example are cycles. Thereby H2 = Z
so long as any two-dimensional chain cycle has the for pQ2, where p is
an integer.
Using this approach we have found the homology groups of a torus
much quicker than by employing the usual simplicial chains and cycles.
We shall give another example. This is a projective plane (Fig. 1.3.4).
Cell homologies We write the action of the boundary operator: a, = QO - QO = 0,
of projective plane aQ2 =2,. Therefore, the integer-valued homology groups are as follows:
Ho = Z, HI = Z2, H2 = O. If we employ the group of coefficients Z2,
there appears a nonzero two-dimensional cycle, the cell Q2. Its boundary
Calculational procedure is equal to 2" i.e. it is equal to zero modulo Z2. Hence, Ho<Rp2, Z2) =
Z2, HI<Rp2,Z2) = Z2, H2(RP2,Z2) = Z2.
As a simple exercise we shall calculate the homology groups in
all the cases considered in the preceding subsection and make sure that
employing the language of cell simplicial chains speeds up substantially
the calculational procedure.
Figure 1.3.5 presents simplicial complexes which determine a Klein
bottle, a sphere with 9 handles and a non-orientable surface of type k. In
all these cases all the vertices of a corresponding polygon are glued into
one point (following the identifications indicated on the polygon sides)
and, therefore, depict a single vertex QO• Thus, each of the polygon sides
(taken separately) represents a one-dimensional cycle on the surface. It
remains to find the boundary of the two-dimensional chains. The only
cell of dimension two is the polygon Q2 itself. We shall calculate the
1.3 General Properties of Simplicial Homology Groups 59

boundary of this cell. In the case of a Klein bottle we have: 8Q2 =


a + {3 - a + (3 = 2{3 f 0; for a sphere with 9 handles we have: 8Q2 =
a)+{3)-a)-{3)+·· ·+ag+{3g-ag-{3g = 0; for a non-orientable surface of
type k we have: 8Q2 = 2a) +...+2ak f o. Accordingly, the homologies of
a Klein bottle are of the form: Ho = Z, HI = Z(a)El1Z2({3), H2 = 0, where Cell homologies
the Abelian cyclic groups generated by a and (3 are denoted respectively of Klein bottle
by Z(a) and Z2((3). The homologies of a sphere with 9 handles are of the
form: Ho = Z, HI = Z(al)El1Z((3I) El1 ... El1 Z(ag) El1 Z((3g) = Z2 g, H2 = Z.
The homologies of a non-orientable surface of type k are: Ho = Z, HI =
Z(al) El1 ... El1 Z(ak-I) El1 Zz(al + ... +ak).

1.3.4 Chain Homotopy. Independence of Simplicial Homologies


of a Polyhedron of the Choice of Triangulation

As has already been noticed, the homology groups are important and for
they do not depend on the choice of triangulation of a polyhedron.

Theorem. Let X be a finite polyhedron and let K), K2 be the sim- Simplicial mapping
plicial complexes corresponding to two arbitrary triangulations of this
polyhedron. Then the simplicial homology groups Hj(KI) and Hj(K2)
are isomorphic and coincide (by definition) with the homology groups Simplicial homologies
Hj(X) of the polyhedron. of the polyhedron

We shall not prove this theorem since the proof is fairly trivial and
technically cumbersome. We shall only give the brief scheme of the proof
displaying its visual geometrical meaning.
Suppose X is a polyhedron on which a certain triangulation gener-
ating a simplicial complex K is fixed. We have introduced before the Chain complex
groups of simplicial chains Cj(K) and the boundary operators 8j . Taking
all these groups together, we may organize them in a sequence of groups
and homomorphisms of the form

o -t Cn(K) ~ ... - t Cj(K) ~ Ci-I(K) - t •••

-t CI(K) ~ CoCK) ~ Z - t 0 .
60 1. Polyhedra. Simplicial Complexes. Homologies

This sequence is called a chain complex (a complex of chains). The


homorphism c is organized as follows: c (E j ajLl~) = Ej aj E ~.
Moreover, we recall that there always holds the identity Oi Oi+ I == 0
for any i and that ker Oi = Zi(K) cycles, Imoi+1 = Bi(K) (boundaries),
Hi(K) = Keroi/Imoi+l.
Suppose we are given two simplicial complexes K and L. The map
f: K - t L is called simplicial if it is continuous and maps linearly each
simplex from K onto a certain simplex from L (i.e. on a simplex of
I the same or lower dimension). In particular, under a simplicial mapping
I
vertices go into vertices. We shall consider a simplicial complex K and
Kx I :I
I its direct product K x I by a unit segment I. As a result we obtain a
I
I "prism" (Fig. 1.3.6). We state that the direct product can be transformed
,,;l'\
,'-.:::.t:-~ into a simplicial complex, i.e. that K x I can be triangulated (generally,
~~--, ............'
K in many ways). We shall point out one of these ways. Clearly, it suffices
to learn to triangulate a direct product Lli x I of one simplex Lli by a
segment I. The triangulation procedure is shown in Fig. 1.3.6 for one-

- and two-dimensional simplexes Lli and Ll2. It is quite clear how this
process should be extended to higher dimensions. Doing so, we have
constructed and shall henceforth consider only such triangulations of the
prism K x I which coincide on the prism bases (Le. on K x 0 and K xl)
with the given triangulation of K, the triangulated prism K x I falling
into the union of prism Lli x I, there Lli runs over all the simplexes of the
complex K. We shall not repeat these conditions each time in the sequel,
but we simply speak of triangulation of the prism K x I compatible with
Fig. 1.3.6 the initial triangulation of K.

Definition. Two simplicial maps f, g : K - t L of the complex K into the


Simplicially complex L are called simplicially homotopic if there exists a simplicial
homotopic mappings map F of the direct product K x I (regarded as a simplicial complex,
where K x I can be triangulated in an arbitrary compatible way, not
Compatible triangulation necessarily the one shown in Fig. 1.3.6) into the complex L, such that
of the prism on the bases of the prism K x I the map F coincides with f and g, i.e.
F(x,O) = f(x), F(x, 1) =g(x).

Simplicial map Let f: K - t L be a simplicial map of two simplicial complexes.


Since it carries each simplicial simplex into a simplex, it is obvious that
the map Ii of simplicial chains Ci(K) of the complex K into simplicial
Chain map chains Ci(L) of the complex L is defined. All the maps fi are homo-
1.3 General Properties of Simplicial Homology Groups 61

morphisms. It is convenient to tabulate this set of homomorphisms and


groups, i.e. to represent them as diagrams, where the identity map Z ...... Z
is denoted bye.

Commutativity
of the diagram

From the definition of boundary operator it follows that this diagram is


commutative, i.e. fi- 18i = 8'di' ec: = c:' 10. The set of maps {Ii} is called
a chain map.
From commutativity of the diagram it is immediate that the ho-
momorphisms Ii induce some homomorphisms li*: Ker8i/Im8i+l ......
KerB:/Im B:+l' i.e. homomorphisms of the homology groups of chain
complexes li*: Hi(K) ...... Hi (L). We shall sometimes write 1* instead
of li*. One can easily verify fulfillment of the following properties.

1) If I: K ...... Land g: L ...... N are simplicial maps of the simplicial Identity self-map
complexes K, L, N, then (g/)* = g*I*.
2) If lK is an identity self-map of the complex K (it is obviously
simplicial), then (lK)*: Hi(K) ...... Hi(K) is an identity homomorphism
(for any i).

Theorem. Let f, g: K ...... L be two simplicially homotopic simplicial


maps of the complex K into the complex L. Then the induced homo- Induced homomorphism
morphisms f* and g* of the homology groups Hj(K)into the homology
groups Hi(L) will coincide for all i.

To prove the theorem, we shall require the concept of chain ho- Chain homotopy
motopy. Let G = {Gi } and G' = {Gn be two chain complexes, i.e.
Gi =Gj(K) and CI =Gi(L). Let 'P and ¢ be two maps of G into G', i.e.
'P = {'Pi}, ¢ = {¢i}, where 'Pi and 'Pi are the homomorphisms Gi ...... CI·
We shall say that we are given a chain homotopy D between the maps
'P and ¢ of the chain complex G into the chain complex G' if we are
given a set D ={Di} of homomorphisms Di : Gi ...... CI+l such that for
each i there holds the equality Di-18i + a:+l Di = 'Pi - ¢i. This can be
written as the following diagram
62 1. Polyhedra. Simplicial Complexes. Homologies

Chain-homotopic maps
of chain complexes

••• - t I - G'j -
G'j +8i+! a;

It is natural to call such two maps <P and 'l/J of chain complexes G and
G' chain-homotopic.
Chain-homotopic maps of chain complexes induce indentical maps
of homology groups. In other words, if <P is chain-homotopic to <p then
all the homomorphisms <P*, 'l/J* : Hj(K) - t Hj(L) coincide.
Indeed, if the chain Z E Gj(X) is a cycle, i.e. aiZ = 0, then by
the definition of chain homotopy we have: <Pi(Z) - 'l/Ji(Z) = Dj-l ai(z) +
~+IDi(Z) = ~+I(Di(Z» = ~+I(w), i.e. under the maps <Pi and'l/Ji the
Chain-boundary images of the cycle Z differ by a chain-boundary, and therefore they are
homologous. This completes the proof.

Lemma. If two simplicial maps I, g: K - t L are simplicially homotopic,


then the maps of the chains <P = {Ii} and 'l/J = {gi} induced by them are
chain-homotopic.

Corollary; If two simplicial maps j, g: K - t L of the complex K into


the complex L are simplicially homotopic, they induce idential maps of
simplicial homology groups Hi(K) - t Hi(L)for all i.

The proof of the corollary follows immediately from the lemma and
from the properties of chain homotopy. It remains to prove the lemma.
Since the maps j and 9 are simplicially homotopic, there exists a
Triangulation of the prism simplicial map of the prism K x I into L, the prism being triangulated in
a compatible way (the triangulation on its bases coincides with the initial
triangulation of the complex K). One should construct a chain homotopy
of the chain complex Gi(K) into the chain complex Gi(L). It suffices to
Chain homotopy specify this chain homotopy only on some individual simplexes of the
complex K. So, let Lli be an arbitrary simplex in K. Since the triangu-
lation of the prism K x I is compatible with (respects) the triangulation
of K, a prism Lli x I embedded into the prism K x I hangs over the Ll j .
The prism Lli x I is triangulated in some way (it is of no importance for
us in what particular way). Accordingly, it is a simplicial complex itself.
1.3 General Properties of Simplicial Homology Groups 63

We now consider all the (i + I)-dimensional simplexes into which the


prism Lli x I is divided, take their sum, denote it by Di(Lli ) and asso-
ciate it with the initial simplex in the prism basis. Since to each simplex
Lli we have assigned a certain chain Di(Lli) there occurs a homomor-
phism Di : Oi - t Oi+I(L). Indeed, it now suffices to take the image of $implicially homotopic maps
the prism Di(Lli) under a simplicial map F. As a result we obtain an induce identical maps
(i + I)-dimensional simplicial chain in the complex L which we denote of homologies
by Di(Lli). It remains to verify the basic property of chain homotopy,
namely, the property that Di-I Oi +O~+I Di = Ii - gi. We shall rewrite this
equality: O~+I = Ii - Di-IOi - gi· Clearly, this relation follows from the
geometrically obvious fact that the overall boundary of the prism Lli x I
consists of three components: 1) the lower base, 2) the side wall of the
prism (which is a cylinder) and 3) the upper case taken with a minus
sign to make allowance for the orientation. This implies the lemma. We
have proved the theorem that simplicially homotopic maps of complexes
induce identical maps of simplicial homology groups.
Indeed, also valid is a stronger (and simpler formulated) statement.
We first recall the concept of continuous homotopy (already not neces- Continuous homotopy
sarily simplicial).
Two maps I and 9 of a polyhedron X into a polyhedron Yare
called homotopic ifthere exists a family of continuous maps 'Pt: X - t Y Homotopic mappings
(depending continuously on the point x E X and on the parameter t),
such that 0 :s:: t :s:: 1 and 'Po = I, 'PI = g. In other words, the map I is
continuously deformed into the map g. Homotopy of the maps I and 9
is denoted by I rv g. It turns out that any continuous map of simplicial
complexes can be approximated by a simplicial map.
Theorem on simplicial approximation. Let I : K - t L be a continuous Continuous deformation
map of a finite simplicial complex K into a finite simplicial complex L. ofa map
Then there exists such a (subdivison) K' ofthe complex K (i.e. the simpli-
cial complex K' obtained.from K by dividing each simplex into smaller
simplexes, for example, using successive barycentric subdivisions) and Successive
such a map I' : K' - t L which is homotopic to the initial map and is barycentric subdivisions
simplicial.
Thus, applying an appropriate comminution (repeated a needed num-
ber of times) to an initial complex one can always deform a continuous
map into a simplicial one. This "bridge" between continuous and simpli-
cial maps is essential in the study of the properties of simplicial homology
groups. We shall not prove this theorem here. See, e.g. ref. [24].
64 1. Polyhedra. Simplicial Complexes. Homologies

Clearly, the concept of continuous homotopy can be formulated


Cylinder in the language of the map of a cylinder (prism). Maps f and 9 are
homotopic if there exists a continuous map F: C x I -+ Y, such
that F(O, x) = f(x) and F(l, x) = g(x). It is clear that for the maps
CPt : X -+ Y it suffices to take the maps CPt (x) = F(t, x), 0 :::; t :::; l.
It turns out that in all the statements formulated and proved above
one can everywhere replace the words "simplicial map" by "continuous
map ". In particular, the continuous homotopic maps f, g: K -+ L induce
identical homomorphisms of the homology groups Hj(K) -+ Hj(L).

Claim. Let K be a simplicial complex corresponding to a fIXed trian-


gulation of a polyhedron X. Let (3i K be a simplicial complex obtained
under a j-stage barycentric subdivion of the initial triangulation (of the
complex K). Then the homology groups of the complex K and of its
j-stage barycentric subdivision coincide (are isomorphic).

The proof (which we omit here) follows from the chain homotopy
Fig.1.3.? property, from Fig. 1.3.7 and from the theorem on simplicial approxima-
tion.
Returning to the theorem on independence of homology groups of
the choice of triangulation, we can note the following. Comminuting
one triangulation (reducing it to smaller fragments), on the one hand
we do not change the homology group (see above). On the other hand,
we cover the complex with increasingly small simplexes. Beginning with
another triangulation, we cover the polyhedron with very small simplexes
Independence of homology also without changing the homology groups. One can prove that the
groups of the choice comminuted cycles of the first triangulation can be associated with some
of triangulation other cycles that combine features of simplicial and cell chains. The same
can be done with another triangulation. From this we deduce isomorphism
of homology groups.
Two polyhedra X and Yare called homotopically equivalent if there
exist continuous maps f: X -+ Y and g: Y -+ X such that their com-
positions f 9 and 9f are respectively homotopic to identity selfmaps I y
Homotopically and Ix. The properties of homology groups immediately imply that ho-
equivalent polyhedra motopically equivalent polyhedra have equivalent homology groups. In-
deed, (jg)* = (ly)* = id,(gf)* = (lx)* = id, i.e. the homomorphisms
f*: Hj(X) -+ Hj(Y) and g*: Hj(Y) -+ Hj(X) are mutually inverse, and
therefore each of them is an isomorphism.
1.3 General Properties of Simplicial Homology Groups 65

1.3.5 Visual Material

Figure 1.3.8 shows schematically the principal geometrical idea of the


theorem on simplicial approximation (see Sect. 1.3.4). The initial continu- Simplicial approximation
ous mapping of polyhedra is subjected to a small continuous deformation
(homotopy), as a result of which the mapping is locally "smoothed out"
and becomes piecewise linear. The process then extend to the entire poly-
hedron. In Fig. 1.3.9 the same idea is realized in a somewhat different
form. The initial continuous map was given a by polynomial of a high Polynomial
degree (see the graph in the centre of the figure). Then the graph was
triangulated, which made it possible to regard the map to become locally
linear (on each simplex) after an appropriate small deformation. We can
see that at the points where the derivative of the map is large ,enough (i.e.
the points located on thin "whiskers" of the surface) the surface should "Whiskers" of the surface
be reduced to especially small fragments. The larger the modulus of the
derivative, the smaller the simplexes. Figure 1.3.10 is a graph of math-
ematical variations of the theme of the appearance of cell homologies
"inside" simplicial ones. Sufficiently small simplexes can be organized
into larger and more complicated figures which themselves can play the
role of elementary objects-cycles. Figure 1.3.11 presents prism triangu-
lation and, therefore, chain homotopy; see the theorem on independence
of homology groups of triangulation. In the upper portion of Fig. 1.3.11
there is an approximation of a continuous map by a simplicial one: in
the centre the polyhedron is cut and there appear nontrivial boundaries.
Two figures drive apart the cut shores. This technique is used below
in the proof of the classification theorem on two-dimensional surfaces.
A surface is cut into triangles of which a flat polygon is glued. Figure
1.3.12 presents schematically a sequence of incidence matrices (rectan- Sequence of
gular tables-plates) of a polyhedron (see Sect. 1.3.1). The product of any incidence matrices
pair of neighbouring matrices is equal to zero. Different forms of chain
homotopy prisms and various ways of their division into elementary
pieces are depicted in Figs. 1.3.13 and 1.3.14.
66 1. Polyhedra. Simplicial Complexes. Homologies

Page 67, Fig. 1.3.8. Theorem on simplicial approximation. From the cy-
cle "Signs of Zodiac" - constellation of Virgo; and simultaneously - the
picture for Bulgakov's novel (Margaret's flight)

Page 68, Fig. 1.3.9. Simplicial approximation of a continuous mapping;


polynomials of a high degree. Medieval German and Scandinavian leg-
ends

Page 69, Fig. 1.3.10. Appearance of cell homologies "inside" simplicial


ones. To M. Bulgakov's '7he Master and Margaret": the giant cat Bege-
mot leaves the place of crucifixion; the dream of Pontius Pilate

Page 70, Fig. 1.3.11. Prism triangulation, chain homotopy and indepen-
dence of homology groups of triangulation. Egyptian legend about star
Sotis; ancient priests in the temple devoted to Sotis

Page 71, Fig. 1.3.12. Incidence matrices of a polyhedron and calculation


of a homology group. Ancient military court; Roman legend; Remus tem-
ple on the top of a pyramid

Page 72, Fig. 1.3.13. Chain homotopy prisms and invariance of homology
groups. To M. Bulgakov's 'The Master and Margaret": crucifixion of Jesus
Christ; non-canonical version of Middle Ages

Page 73, Fig. 1.3.14. Polyhedra and simplicial homologies. To M. Bul-


gakov's '7he Master and Margaret": medieval version of the legend on
crucifixion
2. Low-Dimensional Manifolds

2.1 Basic Concepts of Differential Geometry

2.1.1 Coordinates in a Region, Transformations of Curvilinear


Coordinates

Modem differential geometry is an independent scientific discipline, ex-


ceedingly branched and connected with numerous applications. Today's
edifice of differential geometry exhibits two basic layers. The one which
was historically the first to appear may be conditionally called local dif-
ferential geometry which usually develops in a region of a Euclidean
space. This is the foundation and the first storeys of the whole edifice.
Then there appeared next storeys which took their shape later than the
first ones. They are referred to as "global differential geometry". Here Global differential geometry
local concepts are closely interwoven with global, topological ones. The
backbone of the whole building is the theory of smooth manifolds. On
lower storeys, it is at first local, i.e. events occur in a small enough
domain of a manifold. Ascending, the theory develops global aspects
and finally, on the top, the modem geometry operates with essentially
nonlocal effects.
We shall start with local concepts. We shall consider a Euclidean
space Rn endowed with an ordinary Cartesian (Euclidean) scalar product Euclidean scalar product
and related to a Euclidean basis consisting of unit and pairwise orthogonal
vectors eI, ... , en. In its final form the idea to specify points of the space
Rn using a set of real numbers xl, ... ,xn which are its coordinates rel-
ative to the basis el, ... ,en, is due to Descartes (1596-1650), although
in a less clear form it is already presents by Viete (1540-1603) and
P.Fermat (1601-1665). The concepts oflatitude and longitude (the ana-
logue of abscissa and ordinate) f~r drawing graphs were however used
76 2. Low-Dimensional Manifolds

by N.Oresme (1323-1382) as far back as the XIV-the century. Fruit-


fulness of this idea was realized to the bottom on the appearance of the
Curvilinear coordinates concept of curvilinear coordinates. Before proceeding to their discussion,
Cartesian coordinates we shall dwell on the concept of Cartesian coordinates. Their importance
is first of all in that they allow algebraic definition and investigation of
curves 'Y in lin. The characteristic properties of Cartesian coordinates is
their rigid connection with the Euclidean scalar product (,) of vectors a
and b which are to be written as (a, b) = ~f=l aibi, where ai and bi are
Cartesian coordinates of the vectors a and b. The concept of curvilinear
coordinates was formed under the pressure of concrete mechanical and
physical problems which required a more sophistcated technique than the
ordinary theory of Cartesian coordinates. In many applied problems we
Central force field encounter trajectories of the motion of material points in some force field
whose explicit representation in Cartesian coordinates is either difficult
or too cumbersone. We shall give a simple example.
Let a material point A with a mass equal to 1 move in a plane
in the central force field. Let the centre of gravity be at a point O.
We shall introduce polar coordinates r and <p on the plane, where r
is the distance of the variable of the point A to the point 0 and <p
is the angle between the radius vector a of this point and the x-axis
Fig.2.1.1 (Fig. 2.1.1). Since the point A moves, the r and <p are functions of time
t. Consider the unit vector er at the point A. It is directed along the
radius vector and a = rer. Differentiation with respect to t is denoted by
a point. Then the motion of a material point in the central force field is
described by the differential equation a = j(r)er , where j is a smooth
function of r for r > 0 and r(t) = la(t)l, the quantity lal being the
Euclidean length of the vector a. We shall conider the case of a field
where the attractive force is inversely proportional to r2, i.e. a Newtonian
attraction field Then the finite (compact) trajectories of motion of a point
are ellipses whose equations are of the form 1 + e cos <p = p/ r, where
excentricity p and e are the parameter and excentricity of the orbit, e being less
of the orbit than unity for finite motions. The centre of attraction is in one of the
Polar coordinates ellipse foci. Here polar coordinates simplify substantially the study of
the trajectory of a point, and thus their use is justified. In solution of
concrete mechanical and physical problems there analogously appeared
other convenient curvilinear coordinate systems, such as spherical (for
example, in the study of the motion of a rigid body with a fixed point in
2.1 Basic Concepts of Differential Geometry 77

a gravitational field), cylindrical (in the study of liquid column stability Open region
and surface tension forces), etc.
The analysis of all such coordinate systems suggests that they are
based on some general idea which admits a simple and independent
formulation.
We consider an arbitrary open region C in ]Rn endowed with Carte-
sian coordinates xl, ... , xn. Assigning each point P from C a set n
of real numbers yl(p), ... , yn(p), we define n functions on the re-
gion C. We shall specify curvilinear coordinates in terms of these func-
tions. Since the position of a point can be uniquely determined by its
Cartesian coordinates, the functions introduced above will be written as
yl = yl (xl, ... ,xn), ... , yn = yn(xl , ... ,xn).
Definition 1. A continuous coordinate system in a region C is a family Continuous
of functions yi = yi(x l , ... , xn) determining a one-to-one and continuous coordinate system
map f of the region C onto a region A in another copy of the Euclidean
space. In other words, f is a homeomorphism of the region C onto the
region A.

Of particular interest for applications is the case where the map f


is smooth, i.e. all the functions yi(xl, .. . ,xn) depend smoothly on the
arguments. The term "smoothness" is understood as follows. We require
that the functions have so many continuous derivatives as is required
by the condition of the problem. For simplicity we assume that we take
functions of the class Coo (an infinite number of continuous derivatives).
Definition 2. The Jacobi matrix of a map f : C ~ A is the function Jacobi matrix
matrix
8 8n)
ay
( ii .........
1
~ ~
df= (ax)
= ........ ·~
8xn ••••••••• 8x n
The determinant of this matrix is called the Jacobian of the map f and Jacobian of the map
is denoted by J(f).
Definition 3. A regular curvilinear coordinate system in a region C is a Regular curvilinear
family of smooth functions yi(xl, ... ,XT) determining a one-to-one map coordinate system
of C onto a region A in ]Rn, the Jacobian of the map f being nonzero
at all points of the region.
78 2. Low-Dimensional Manifolds

Implicit function theorem The implicit function theorem implies that in this case the map i-I,
inverse to f, is also smooth. Thus, a regular curvilinear coordinate system
is given by a smooth one-to-one map which establishes, in particular, a
homeomorphism between the regions C and A. Such maps are called
Diffeomorphism diffeomorphisms. Suppose that in a region we are given a set of smooth
functions yl(p), ... , yn(p). How shall we know whether it specifies a
Regular transformation regular coordinate system in the region?
of coordinates
Lemma 1. Suppose the Jacobian of a family of smooth functions yl (P),
... ,yn(p) is nonzero at all points of the region. Then for each point
there exists an open neighbourhood in which these functions determine
a regular (localj coordinate system.
At the same time this family offunctions may fail to determine the
global curvilinear coordinates in the entire region.

The proof of the first statement follows from the implicit function
r
theorem. The second statement means that the map 1 of the region A
onto the region C may be not existent. We shall present a simple example.
Consider as regions C and A a plane ]R2 with the origin punctured out.
The position of the point P on C will be given by a complex number
z = rei"'. The map f will be given as f(z) = z2 (Fig.2.1.2). Clearly,
the Jacobian of this map is no'nZero (on the entire region), but f is not
a one-to-one map and, therefore, does not have an inverse single-valued
FiQ.2.1.2 map.
We shall define a regular change of coordinates. Suppose in C we
are given two regular coordinate systems, i.e. two diffeomorphisms f:
C - t A c ]Rn(yl, ... ,yn) and g: C - t B c ]Rn(zl, ... , zn) are fixed.
Since these maps are one-to-one, it follows that the map assigning to the
coordinates yi(p) of the point P its coordinates zi(p) is well defined.
There appears a map gr 1 = q : A - t B (Fig.2.1.3). As a result, we
obtain the functions zi = zi(yl, ... ,yn) which determine this transfor-
mation. We call this transformation a regular change of coordinates in
a region. Our consideration is based on the following statement.

Lemma 2. The map zi = zi(yl, ... , yn) is a one-to-one and smooth map
of a region A onto a region B with a nonzero Jacobian.
2.1 Basic Concepts of Differential Geometry 79

The proof follows from the properties of the Jacobi matrix. Recall
that the Jacobi matrix of a composition of maps falls into the product of
the Jacobi matrices of these maps.
The system of curvilinear regular coordinates yl, . . . ,yn in a region
C determines the families of coordinate lines in this region. Namely, the Coordinate lines
i-th coordinate line passing through a point Po is a smooth trajectory de-
scribed by the set of equations yj (P) = d, j =!= i and 1 ~ j ~ n, where d
are some fixed constants, namely, d = yj(Po). In other words, along the
i-th coordinate line there changes only the coordinate yi, the rest of the
coordinates being fixed. Thus,jrom each point there goes n different co-
ordinate lines (Fig. 2.1.4). We shall give examples of classical curvilinear Fig.2.1.4
coordinates on the plane and in I? We determine Cartesian coordinates

~o.~.
x, y on the plane. Taking as a region C a plane with a discarded ray
(x ~ 0, y = 0), we can define the functions <p = arctan y/x, r = Jx2 +y2
on this plane. This change of coordinates is also given by the formulas
x = r cos <p, y = r sin <po A rectangular net of Cartesian coordinates be- Fig.2.1 .5
comes a polar net of coordinate lines (Fig. 2.1.5). Polar coordinates
Cylindrical coordinates r, <p, t in R3 are well defined, for example, Cylindrical coordinates
in a region obtained by discarding a half plane (x ~ 0, y = 0) from a
three-dimensional space. The formulas for the change of coordinates are
as follows: x = r cos <p, y = r sin <p, z = t (Fig. 2.1.6).
Spherical coordinates r, e, <p in R3 are well defined, for example, Spherical coordinates
in a region obtained by discarding from the space of the same half plane

~-r
as in the preceding example. The formulas for the change of coordinates
e e e
are: x = r sin cos <p, y = r sin sin <p, z = r cos (Fig. 2.1.7).

Fig. 2.1.6

2.1.2 The Concept of a Manifold. Smooth Manifolds. Submanifolds


and Ways of Defining Them. Manifolds with Boundary. Tangent
Space and Tangent Bundle

Fig.2.1 .7
There presently exist several ways of drawing maps of earth's surface
or its separate regions. All of them are in any case reduced to one and
the same procedure, namely, to projecting a convex spherical surface of
the globe (or its parts) onto a plane. It is more or less obvious that it
is impossible to make a one-to-one continuous projection of the whole
80 2. Low-Dimensional Manifolds

sphere onto a region of a plane (or onto a whole plane). This assertion can
Medieval cartographers be given a rigorous mathematical proof. Medieval cartographers did not
however realized this essential fact immediately. Superficial experience
of an observer allows him to hypothesize that earth's surface is flat. So,
the earth was thought of as flat and limited by the "boundary of the
world". The first medieval maps of the world were drawn on a plane,
accordingly, in the form of a disc or a rectangle where cartographers
placed the whole world known at that time. See, for instance, medieval
maps dated traditionally X-th and V-th centuries A.D. (Fig. 2.1.8). The
same are also the maps, for example, of the .X-XV-the centuries A.D.
The same is the world-known map of 1527 made by a Portuguese Diogo
Ribeiro [34].
After the discovery of the spherical shape of the earth, after the trips
of Christopher Colomb (1492), Vasko de Gama (1497-1498), Magellan
(1515-1592), the necessity of maps taking into account earth's sphericity
was finally realized. The first attempts were pictures of a ball, a globe
turned to the viewer with one side which was most important for the
cartographer. These pictures represented corresponding regions which
were, however, gradually distorted with approaching the visible boundary
of the ball. Such is, for example, the map of the world due to Stabius-
Diirer (1515) [34]. At the same time, cartographers were not satisfied with
this way of mapping since the invisible part of the ball required a separate
Fig.2.1.8 representation; moreover, the difficulties of representing a convex globe
surface in a flat picture needed engraver's and draftsman' skill. Therefore,
various ways of projecting large regions of earth's surface onto a plane
with allowance made for inevitable distortions were invented already in
the XVI-th century. Great contribution to the development of the theory
of projection and its applications (to cartography, painting and drawing)
was made, in particular by A. Diirer (1471-1528). Thus, beginning with
the XVI-th century cartographers use various tricks cutting a sphere into
several pieces each of which is then separetely projected onto a part
of a plane. The initial sphere (globe) is restored from them by way
of glueings according to the rules indicated on the flat map. The most
frequently used technique is to cover the surface of the globe with two
regions each of which is a bit larger than a hemisphere and then to project
Fig.2.1.9
them separately onto a plane (Fig. 2.1.9). Glueing to two discs obtained
along their common part (along a narrow ring), we can restore the initial
2.1 Basic Concepts of Differential Geometry 81

picture on the globe. The map of the world can be glued of several pieces
(Fig. 2.1.9). Thus, a sufficiently complic"ated object (a sphere) is obtained
of several simpler objects (of two or more discs) by way of glueing along
some common part of theirs.
This idea underlies the construction of a wide class of geometrical
objects called manifolds. Manifold
The concept of a manifold acquired its most clear form due Gauss Gauss
(1777-1855) when he was engaged in the studies in the field of geodesy
and carthography of earth's surface. In a practical drawing of maps of
rather large regions of earth's surface, they are divided into smaller, par-
tially overlapping regions, each assigned to a separate group of experts.
They drawn a map of each separate region endowed with reference points
(the names of villages, rivers, etc.). In compiling the general atlas, these
separate maps are sewed, glued along those regions which overlapped,
and thus they are represented in several local maps. Linking separate
local maps is realized by comparison and superposition of their common
reference points.
This is just the idea that underlies the modem concept of a mani-
fold. The term "manifold" was introduced to mathematics by B. Riemann Riemann
(1826-1866) in his famous lecture "On the hypotheses underlying geom-
etry". Roughly speaking, manifolds are geometrical objects obtained by
glueing open discs (balls) like a papier-mache is glued of small paper
scraps. To this end, one first prepares a clay or plastecine figure which is
then covered with several sheets of paper scraps glued onto one another.
After the plasticine is removed, there remains a two-dimensional surface.
We now give the mathematical definition.

Definition 4. A topological Hausdorff space M is called an n-dimensional


topological manifold if it is represented as a union of its open subsets Topological manifold
Ui and if for each of these subsets a homeomorphism 'Pi : Ui -+ Dn is
fixed which maps Ui onto a standard open disc in a Euclidean space Rn.
The homeomorphism 'Pi is called a coordinate map, the pair (Ui, 'Pi) is Coordinate map
called a map, or a chart (or a local chart), the union of charts {(Ui .'Pi)} Local chart
is called the atlas on the manifold M. The number n is the dimension of Atlas
the manifold. The manifold is called compact or noncompact according Dimension of the manifold
as the corresponding topological space is compact or noncompact. Compact and
noncompact manifolds
82 2. Low-Dimensional Manifolds

Each coordinate map determines coordinates on the set Uj since


this map determines on the chart the family of continuous functions
Xl (P), ... , xn(p} which can be regarded as coordinates of a variable
point P (Fig.2.1.10). With each atlas on a manifold there is naturally
associated the concept of transition functions. We shall consider two
;btl I~ ~ \ arbitrary charts Uj, Uj and their intersection Uj n Uj. On this intersection

~~~,
two coordinate maps 'Pj : Uj n Uj -+ 'Pj(Ui n Uj} c Dn and 'Pj : Ui n
Uj -+ 'Pj(Uj n Uj} c Dn are defmed (Fig.2.1.10). Since a composition
of homeomorphisms is a homeomorphism, on the open subset 'Pi(Ui n
IRn IR tl Uj} in the disc Dn the homeomorphism 'Pij = 'Pj'Pjl mapping the set
Fig.2.1.10 'Pi(Ui n Uj} onto the set 'Pj(Ui n Uj} is well defined.

Definition 5. The homeomorphisms 'Pij are called glueing functions or


Transition functions transition functions (in a given atlas).

Determination of transition functions makes it possible to restore


the whole manifold if individual charts and coordinate maps are al-
ready given. Glueing functions may belong to different functional classes,
which makes it possible to specify within a certain class of topologi-
Smooth and cal manifolds more narrow classes of smooth, analytic, etc. manifolds.
analytic manifolds We shall introduce Cartesian coordinates on the disc Dn. Then the
homeomorphism 'Pij can be written as a family of functions yk
li~(xl , ... ,xn}.

Definition 6. A manifold M is called smooth, of class Cr , or analytic if


all the functions li~ are respectively smooth, of class Cr, or analytic.

Piecewise smooth, Lipschitz and other types of manifolds are defined


in a similar way.
Two topological manifolds are called homeomorphic if so are the
Homeomorphic and corresponding topological spaces. Smooth manifolds are called diffeo-
diffeomorphic manifolds morphic if they are homeomorphic and if the homeomorphisms I and
I-I are smooth maps.
Examples The simplest examples of manifolds are: open regions in a Euclidean
of manifolds space, graphs of smooth functions of the form xn+1 = f(x l , ••• , xn} in
jRn+ I. One of the important ways of determining manifolds is as follows.
We consider in jRn a family of smooth functions 9i(X I, ... , xn }, 1 ::; i ::;
2.1 Basic Concepts of Differential Geometry 83

k ::; n. Then the common level surface, at each of whose points the
rank of the Jacobian matrix (a9/ax) of this family is equal to k, i.e. is
maximal, is a smooth manifold of dimension n - k. This follows from

°
the implicit function theorem. Thus, manifolds can be given as sets of
solutions of systems of equations 9i(X I, ... , xn) = in ]Rn. For k = 1 System of equations
we obtain a hypersurface. Smooth manifolds are, for example, standard for the manifold
spheres sn, real and complex projective space ]Rpn and cpn, groups of
orthogonal matrices SO(n), unitary matrices U(n), special unitary matri-
ces SU(n), symplectic matrices Sp(n), etc.
We have naturally come up to the concept of a submanifold. Hence- Submanifold
forth we mainly deal with smooth manifolds, the therefore will not each
time repeat the smoothness condition.

Definition 7. A subset V in a manifold Mn is called a smooth sub-


manifold of dimension n - k if for each point P E V there exists an
open neighbourhood U(P) with local coordinates xl, ... , xn and a set
of k smooth functions 91, ... ,9k defined on U(P), such that the rank
(a9/aX) = k and V n U(P) = {x E U(P) : 9i(X) = 0,1 ::; i ::; k}. In
other words, the intersection of V with the neighbourhood U(P) must
coincide with the common level surface of the functions 91, ... ,9k.

The condition on the rank of the Jacobi matrix implies that all the Rank of the
functions 91, ... ,9k are fUnctionally independent in the given 'neighbour- Jacobian matrix
hood. Independent functions

Definition 8. A subset V in a manifold M n is called a smooth subman- Two definitions


ifold of dimension n - k if for each point P E V there exists an open of a submanifold
neighbourhood U(P) in which one can introduce local curvilinear coor-
dinates xl, ... , xn such that VnU(p) = {x E U(P) : xi = 0,1 ::; i ::; k}.

Proposition 1. Definitions 7 and 8 are equivalent.

Let us pay attention to the fact that this defmition of a manifold does
not embrace a number of natural geometrical objects, for instance," an n-
dimensional closed disc. The point is that for points located on the disc
boundary there exists no open neighbourhood homeomorphic to an open
disc in ]Rn. This category of objects is within the scope of the concept
of a manifold with boundary.
84 2. Low-Dimensional Manifolds

Definition 9. Let I(x) be a smooth function given on a smooth manifold


Mn (in the sense of Definitions 4 and 6). A closed subset vn defined by
Smooth manifold the inequality I(x) ~ o(or I(x) ~ 0) is called an n-dimensional smooth
manifold with boundary provided that its boundary 8V specified by the
equation I(x) = 0 is a nonsingular smooth submanifold of dimension
n - 1 in Mn (Fig. 2.1.11). A compact manifold without boundary is
called closed.

Points of a manifold with a non-empty boundary fall into two classes:


F·Ig...
2 1 11 interior points (centres of neighbourhoods homeomorphic to an open
disc) and boundary points having a neighbourhood homeomorphic to
half a disc. A boundary point lies on the equatorial cross-section of the
disc (Fig. 2.1.12).
With each point x on the smooth manifold Mn there is naturally
associated a linear n-dimensional space called tangent.

Tangent vector Definition 10. We shall say that a point x E M a tangent vector a is
given in case in each local regular coordinate system a set of numbers
aI, ... , an (coordinates of the vector) is given which is formed under the
coordinate change (x) --? (x') by the lawai' = Ei(8xi' /8x i )ai, where
8xil /8x i are coefficients of the Jacobi matrix of the coordinate change.

Definition 11. We shall say that at a point x E M we are given a tangent


vector a if the first-order linear differential operation D = Ei ai (8/8x i )
with constant coefficients is specified on the functions defined in the
neighbourhood of the point x.

These two definitions are equivalent. The correspondence between


them is established as follows. If at some coordinate system at a point Xo
we are given a tangent vector a = (a l , ••. ,an), then the differentiation of
Derivative in direction smooth functions I with respect to the direction of this vector is uniquely
of the vector defined, namely,
2.1 Basic Concepts of Differential Geometry 85

Proposition 2. The set of all tangent vectors to a manifold Mn at a


certain point x forms a linear space of dimension n. This space is called Tangent space
a tangent space to the manifold at a given point and is denoted by TxM. to the manifold

If in the neighbourhood of a given point a coordinate system


xl, ... ,xn is fixed, then at this point there naturally arise n linearly
independent tangent vectors ei = 8/8xi (i.e. differential operators). They
correspond to differentiations along the coordinate lines passing through
the point x. From regular character of local coordinates it follows that
these vectors form a basis in the tangent space TxM. Basis in the
Since the tangent space is defined at each point of the manifold, there tangent space
arises a natural geometrical object which is a manifold of dimension 2n.

Definition 12. A tangent bundle T*M is the totality of all pairs of the Tangent bundle
form (x, a), where x is a point of the manifold M and a is the tangent
vector to the manifold at this point.

To verify that T*M is a 2n-dimensional smooth manifold, it suf-


fices to introduce regular local coordinates on T*M. Let xl, ... , xn be Local coordinates
coordinates in the region U on M. Then at each point x E U there arises
a basis in TxM composed of vectors 8/8xi , and each tangent vector
a E TxM is decomposed by this basis: a = Ei ai (8/8x i ). Pairs of the
form (x, a), where x E U, form a region W in T*M. Local coordinates
in W have the form x I , ... , xn, aI, ... , an. It can be directly verified
that the transition functions are smooth, and the assertion follows.

2.1.3 Orientabllity and Non-Orientabllity. The Differential of a


Mapping. Regular Values and Regular Points. Embeddings and
Immersions of Manifolds. Critical Points of Smooth Functions on
Manifolds. Index of Nondegenerate Critical Points and Morse
Functions

Definition 13. A manifold M is called orientable if on it there exists Orientable manifold


an atlas (Ui, "Pi) such that the lacobians of all the transition functions
"Pij from chart to chart are positive for all intersecting pairs of regions.
A manifold is called oriented if such an atlas on it is given and fixed. Oriented atlas
86 2. Low-Dimensional Manifolds

Non-orientable manifold Manifolds that do not satisfy this property are called non-orientable. This
means that there always exists apair of charts on whose intersection the
Jacobian of the transition function is negative.

There also exists another version of definition of orientability. Let


us consider an arbitrary point x on the manifold M and fix in the tangent
Reference frame plane TxM the reference frame e(x) which is the basis. Let ,(t) be an
arbitrary smooth closed path which is contained in the manifold, starts
and ends at the point x (i.e. ,(t) is a loop). Recall that a smooth path
is a smooth map, of a segment [0,1] into the manifold, the velocity
vector 'Y(t) of the curve ,(t) being nonzero at all points of the curve.
Since each point y of the path, is covered by a chart in M, one can
always determine a continuous nondegenerate deformation of any refer-
ence frame given at this point into a close point along the path,. To do
so, it suffices to transport the point y and the path, into a Euclidean disc
using the coordinate mapping and to realize any continuous nondegener-
ate deformation of the reference frame along the path in the ball, which
is already not at all difficult. Clearly, the continuous frame deformation
along the path is not uniquely defined. Let eC/(t)) be a tangent frame
obtained through a nondenerate deformation of the tangent frame e(x)
Fig.2.1.13 along the path, from the point x to the point ,(t) (Fig. 2.1.13.)

Definition 14. We shall say that a manifold M is orientable if for any


closed path, starting and ending at a point x any continuous nondegen-
Deformation of erate deformation of the tangent frame e(x) along the path, transforms
the tangent frame it to the tangent frame e'(x) in the tangent plane TxM which (the frame)
and orientability has the same orientation as the initial one. This means that these two
frames differ from each other by a transformation with positive determi-
nant. Otherwise, the manifold is called non-orientable, i.e. there exists
such a closed path starting and ending at the point x and such a con-
tinuous nondegenerate deformation of the frame along the path that the
frame returns to the point x with a reverse orientation.

Proposition 3. Definitions 13 and 14 are equivalent.

Tangent bundle Lemma 3. A tangent bundle of any smooth manifold is an orientable


is always orientable manifold.
2.1 Basic Concepts of Differential Geometry 87

Distinctions in the properties of orientable and non-orientable man-


ifolds can be easily seen in the two-dimensional case. Two-dimensional
manifolds admit a simple classification. To formulate the result, we shall
recall two operations: a) glueing of a handle, b) glueing of a Mobius
strip. A handle is the direct product of a segment by a circle, i.e. a ring,
a cylinder (Fig. 2.1.14a). To glue a handle to a two-dimensional manifold,
one should do the following. One should discard two sufficiently small
non-intersecting discs from the manifold (Fig. 2.1.14) and glue to the two
holes a cylinder identifying its boundaries with those of the holes. To
perform the second operation, we require a Mobius strip (Fig. 2.1.14b)
obtained from a rectangle through identification of its opposite sides with
orientation reversal (Fig. 2.1.14). The Mobius strip can also be obtained
from a usual flat ring through identification of diametrically opposite
points on one of its boundary circles. Then we discard from a manifold
a sufficiently small disc and identify the boundary of the hole obtained
Fig. 2.1.14
with that of the Mobius strip.

Theorem 1. (Classification theorem of two-dimensional closed sur- Classification theorem


faces). Any two-dimensional smooth compact connected closed manifold for 2-surfaces
is homeomorphic (and diffeomorphic) to one of the following manifolds:
a) a two-dimensional sphere with 9 handles (Fig. 2.1.15), where 9
is a non-negative integer (the genus of the surface),
b) a two-dimensional sphere with k Mobius strips, where k is a
positive integer.
Type 1 manifolds are orientable and type 2 manifolds are not. Man-
Fig.2.1.15
ifolds of type 1 are not homeomorphic to those of type 2. Within each
of these two series, manifolds corresponding to different 9 or k are not
homeomorphic either.

Non-orientability of manifolds of series b) is seen from Fig. 2.1.16.


A path 'Y which once traverses the Mobius strip along its axis has the
property that the two-dimensional tangent frame deformed continuously
along the path returns to the initial point with a reverse orientation.
A great role in geometry is played by the concept of the differential
of a smooth map of one manifold onto another.
Fig. 2.1.16 ~
88 2. Low-Dimensional Manifolds

Definition 15. Let f : Mn -7 NP be a smooth map of smooth manifolds.


Differential The differential df of the map f at a point x E M is such a linear map of
of the map a tangent space TxM into a tangent space TyN, where y = f(x), which
is defined in local coordinates xl, ... ,xn on M and yl , ... ,yP on N by
the Jacobi matrix of the map f, i.e. (df(a»i = 'Li8yi/8xi)ai .

The differential of mapping admits several different interpretations.


We present here the interpretations especially instructive for what fol-
lows. We consider on M a smooth curve ')'(t) passing through a point
x. Let x = ')'(0). In local coordinates xl, ... ,xn the curve is given by
the family of functions xl(t), ... , xn(t). For any tangent vector a to M,
at the point x there always exists a smooth curve ')' which has the vec-
Velocity vector tor a = (ai, ... , an) as its velocity vector at the point x, i.e. i'(0) = a.
From now on we denote the differentiation with respect to time by a
point. Under the map f the curve ')' goes into a certain smooth curve
g(t) = f(')'(t», in a manifold N, which passes through the point y = f(x).
At the point y this curve has some velocity vector b = g(O).

Lemma 4. There holds the equality b = df(a).

The proof is immediate from the rule of differentiation of composite


functions.

7i/Jf
Thus, the differential of mapping can be defined in another way. We

1J7
take a tangent vector a at a point x E M and any smooth curve ')' which

1iID f!;fj
has a as its velocity vector at the point x. We consider its image in the
I- manifold N and calculate the velocity vector of the curve iT at the point
y = f(x). As we have seen, the vector obtained is just the image of the
Fig.2.1.17 vector a under the mapping df: TxM -7 TyN (Fig.2.1.17).

Definition 16. Let f : M -7 N be a smooth map of manifolds. Then


x E M is called a regular point of the map f if the differential df at the
Regular points point x is an epimorphism, i.e. if it maps the space TxM onto the entire
of the mapping space TyN, where y = f(x). Then yEN is called a regular value of the
map f (or under the map f) if all of its pre-images are regular points in
M. If f-l(y) = 0, the point y will also be regarded (by defmition) as a
Regular value regular value.
2.1 Basic Concepts of Differential Geometry 89

The property of being a regular point or a regular value does not


depend on the choice of local coordinates on M and on N.
Proposition 4. Let f : M --+ N be a smooth map of manifolds and let
Pre-image of the
yEN be a regular value of the map f. Then its pre-image P = rl(y)
regular value
(including the case when it is non-empty) is a smooth submanifold, in
M, of dimension dim M - dim N.

Definition 17. Let f : M --+ N be a smooth map. It is called an immer- Immersion


sion if at each point x E M the differential df is a monomorphism, i.e.
if there exists a zero kernel. If, besides, f is a one-to-one map of the
manifold M onto its image f(M) in N and f(M) is a closed subset in
N, then the map f is called an embedding. For topological manifolds Embedding
M the concept of embedding is defined as a homeomorphism of the
manifold M onto the closed subset N.

The image of the smooth manifold M, when embedded in N, is


a smooth submanifold in N in the sense of Definitions 7, 8. The role
of immersions and embeddings in geometry can be explained on the
following example.
We have already defined above a smooth manifold in the language of Realization
charts and atlases irrespective of its realization in the form of a surface in of the manifold
a Euclidean space. On the other hand, for investigation of many properties in Euclidean space
of manifolds it is useful to represent them in the form of a surface in
]R.q. The class of submanifolds in finite-dimensional Euclidean spaces
appears to coincide with the class of abstractly given manifolds.
Theorem 2. Any finite-dimensional manifold M of smoothness class Cr ,
where 1 ::; r ::; 00, both compact and noncompact, admits embedding
of the class Cr in a certain finite-dimensional Euclidean space of a
sufficiently high dimension.
Stronger assertions actually hold

Theorem 3. (The Whitney theorem). Any n-dimensional manifold of Whitney theorem


smoothness class Cr, where 1 ::; r ::; 00, both compact and noncompact,
admits for n ~ 1 embedding of the class Cr into a Euclidean space ]R.2n
and for n ~ 2 immersion of the class Cr into a Euclidean space ]R.2n-l.
90 2. Low-Dimensional Manifolds

The proof of this fact is rather nontrivial (on the contrary, Theo-
rem 2 is proved rather simply using partition of unity). For details see,
for example, Refs. [27], [31], [35]. In a simpler way, on the basis of the
Sard theorem Sard theorem one establishes a weaker theorem, namely, that any n-
dimensional manifold, both compact and noncompact (again of the class
or, where 1 ~ r ~ (0) admits embedding of the class or into 1R2n+1 and
Weak Whitney theorem immersion into 1R2n. This is the so-called weak Whitney theorem, as dis-
tinct from the strong Whitney theorem (Theorem 3). In some particular
cases Theorem 3 can be strengthened. For example, if n is positive and is
not the power of 2, then any smooth manifold Mn admits a smooth em-
bedding in 1R2n- 1• On the other hand, for any n = 2Q, there q ~ g, there
exist closed smooth n-dimensional manifolds admitting neither smooth
Embeddings nor topological embeddings into 1R2n - 1• This refers, for example, to a
and immersions real projective space ]Rpn, where n = 2Q• For example, a projective plane
of projective space m>2 is not embedded into 1R3 even topologically, but is immersed into
it. For details see, for example, Ref. [31].
Therefore, without loss of generality one can further on consider
smooth manifolds realized in the form of smooth surfaces in a finite-
dimensional Euclidean space.
Critical points Let f : Mn -+ N m be a smooth map. We shall say that a point
x E M is critical for the map f if the differential df : TxM -+ Tf(x)N
has a rank lower than m = dim N. In this case we shall call a point
Critical values y = f(x) the critical value o/the map f. According to the Sard theorem
(see, e.g. Ref [27]) the set of all critical values of the smooth map f :
M -+ N always has measure zero in the manifolds N. We shall consider
a particular case where N = lRl is a real straight line. Since dim Tf(x)N =
1, the point x is critical for f if and only if dflx = O. The critical
points x of the smooth scalar function f on M are found from the
system of equations {)f / {)x i = 0, 1 ~ i ~ n, i.e. grad f = O. We
Nondegenerate shall say that the critical point x of the smooth function f is called
critical points nondegenerate if the matrix of the second differential (& f /{)xi{)x j ) is
nondegenerate at this point. Since grad f(x) =0, this critical point is well
defined, i.e. independent of the choice of local coordinates. The function
Morse function f on the manifold M is called a Morse function if all of its critical
points are nondegenerate. The second differential d2f can be treated as
a symmetric bilinear form on the tangent space TxM. Let a, bE TxM.
We include the vectors a and b into smooth local vector fields A and
2.1 Basic Concepts of Differential Geometry 91

B respectively and put d2/(a, b) = A(B(f))lx. Clearly, the form ~ I


thus defined is symmetric and its matrix relative to the basis 818xi has
the form &1 18xi8xi. This form is called the Hessian of the junction.
The index of the nondegenerate critical point x for the function I is the Index of the
maximal dimension of the subspace V c TxM on which the Hessian d2I nondegenerate
is negative definite. In other words, the index is the number of negative critical point
squares after reduction of the form d2I to a diagonal form. It turns out
that in the neighbourhood of a nondegenerate critical point of index A
there always exist local regular coordinates xl, ... ,xn such that relative
to these coordinates the function I is written in the neighbourhood in the
following form (the so-called Morse lemma, see e.g. Ref. [35]): Morse lemma

Thus, the function is reduced to the canonical form not only at one point,
but simultaneously in a whole neighbourhood. Do Morse functions exist
on manifolds? The answer is given by the following theorem (see e.g.
Ref. [35]).
1) On any smooth compact manifold there exist Morse functions. Existence of
2) Morse junctions are everywhere dense in the space ofall smooth Morse functions
junctions on the manifold.
3) Any Morse junction has only a finite number of critical points
on a compact manifold.
4) There exists an everywhere dense subset S in the set of all
Morse junctions, such that to each critical value of any function I E S
there corresponds only one critical point.
In this sense the Morse functions are "typical" (are functions "in Morse functions
general position") in the space of all smooth functions. are functions in
general position

2.1.4 Vector and Covector Fields. Integral Trajectories. Vector


Field Commutators. The Lie Algebra of Vector Fields on a
Manifold

Many physical and chemical processes are determined by vector fields


on manifolds. For example, a flow of an ideal incompressible liquid is Flow of fluid
modelled by a nondivergent vector field.
92 2. Low-Dimensional Manifolds

Vector field Definition 18. We shall say that on a smooth manifold M a smooth vector
field is given if at each point x E M a tangent vector a(x) smoothly
depending on the point is given.

If xl, ... , xn are local coordinates, then the vector field in the neigh-
bourhood of some point can be written in the form of a family of
functions, i.e. coordinates of the point a, namely: al(x l , ... , xn), ... ,
Vector field as ordinary an(xl., ... , xn). Each vector field determines in a one-to-one manner
differential equation a system of ordinary first-order differential equations on the manifold,
nameIy.. x.i -- ai(x I , ... ,xn) , 1 <
_ z. <
_ n.
Along with the tangent space TxM, at each point of the smooth
Cotangent space manifold a cotangent space T; M is defmed, which is a dual (conju-
gate) space to TxM. The elements of the cotangent space (which we
shall sometimes call covectors) are linear functionals defined on TxM.
Covectors Clearly, the tangent and cotangent spaces are linearly isomorphic and
therefore are of the same dimension. By analogy with Definition 18 we
Covector field shall say that a smooth covector field is given on M if each point x
a covector, €(x) smoothly depending on the point is given. By analogy
Cotangent bundle with tangent bundle we construct a cotangent bundle T* M whose points
are pairs (x,O, where x E M,€ E T;M. The space T*M is a smooth
2n-dimensional manifold homeomorphic to the manifold T*M.

Integral trajectory Definition 19. The trajectory ')'(t) on the manifold M is called the inte-
gral trajectory of the vector field a(x) if and only if its velocity vector
coincides at each point with the field vector a, i.e. 7(t) = a(")'(t».

The integral trajectories of the field a are solutions of the system of


Vector field, as linear differential equations x(t) = a(x(t». Each vector field a can be treated as
differential operator a linear differential operator with variable coefficients on the manifold.
Indeed, if a is a field and ')' is its integral trajectory (recall that for
sufficiently small t the equation x(t) = a(x(t» always has solutions in
case the righthand is smooth), then for any smooth function f on the
manifold one can determine its derivative along the vector field a. To do
so, one should examine at each point the following expression:

a(f) = lim f(")'(t» - f(x) where x =')'(0) 7(0) = a/x) ,


t ....o t '
2.1 Basic Concepts of Differential Geometry 93

i.e. 'Y is the integral trajectory of the field which goes from the point
x. Sometimes the notation df Ida is used rather than a(f). Clearly, the
function a(f) can be written in the form

afax i
a(f) = '" -. -
~axl dt
= ( '" ai -a. ) f
~ ax l

Thus, when written relative to the basis ei = a I ax i , each vector field a


can be interpreted as a linear differential operator.

Definition 20. Let a and b be two vector fields. Then the vector [a, b] ==
ab - ba called the commutator (or the bracket) of the fields a and b is Commutator
uniquely defined. Here ab and ba are compositions of the differential of vector fields
operators a and b.

In local coordinates xl, ... , xn the field [a, b] has the following
components:

Although the differential operators ab and ba have order two, their dif-
ference is a first-order operator, which just means that the space of all
smooth vector fields on the manifold is closed under the commutation
operation (the vector field commutator is again a vector field). We shall
give a geometric interpretation of the vector field commutator. Let a and Geometric interpretation
b be vector fields defined in the neighbourhood of a point x. It turns of the local commutator
out that one can construct a curve 'Y on M, whose velocity vector at the
point x = 'Y(O) coincides with the values of the field [a, b] at the point
x. Let t be a sufficiently small number and let 'YI be an integral curve
of the field a going from the point x. Let 'Y2 be an integral curve of the
field b going from the point 'YI (t); then let 'Y3 be an integral trajectory
of the field -a going from the point 'Y2(t); finally, let 'Y4 be an integral
curve of the field -b going from the point 'Y3(t). We now specify the
curve 'Y, going from the point x, by the equality 'Y(t2) = 'Y4(t). Then
one can verify that [a, b]f(x) = -y(O)f for any smooth function f. Thus,
94 2. Low-Dimensional Manifolds

the commutator of vector fields measures geometrically the degree of


openess of the tetragon obtained by a successive displacement by one
and the same quantity along the fields a, b, -a, -b. If the fields a and
b commute, the tetragon is closed. In particular, for commuting fields
a and b, the displacements from the point x by the quantity c in the
direction of the field a and then by the quantity 'T] in the direction of the
field b lead us to the same point as the displacement by the quantity 'T]
in the b direction and then by the quantity c in the a direction.

Proposition 5. If two vector fields a and b are tangent to a smooth


submanifold V at each of its points, then their commutator is also tangent
to this submanifold.

The set of vector fields on a manifold is naturally associated with


one of the most important concepts of differential geometry - the concept
Lie algebra of a Lie algebra.

Definition 21. A linear (either finite- or infinite-dimensional) space G is


Commutator said to be a Lie algebra if in this space a bilinear operation [,] is given
in Lie algebra called the commutator and meeting the requirements:
1) the operation is skew-symmetric, i.e. [a, b] = -[b, a] for any
vectors a, b, E G,
Jacobi identity 2) the operation satisfies the Jacobi identity [a, [b, c]] + [c, [a, b]] +
[b, [c, a]]= 0 for any a, b, c E G.

Each Lie algebra is naturally associated with an important operation


Adjoint representation called an adjoint representation of the Lie algebra. For each a E G we
set by definition ada(b) = [a, b]. Clearly, ada is a linear operator on the
Lie algebra.

Differentiation Lemma 5. The operator ada is differentiation of the Lie algebra, i.e.
of the Lie algebra satisfies the Leibniz formula:
and the Leibniz formula
ada[b, c] = [adab, c] + [b, adac]

We shall present the simplest examples of Lie algebras. The alge-


bra of all linear operators of a linear space is a Lie algebra under the
commutation operation given by the formula [A, B] = AB - BA. The
2.1 Basic Concepts of Differential Geometry 95

space of all square matrices of order n is also a Lie algebra. Of great


importance for various applications is the following infinite-dimensional Infinite-dimensional
Lie algebra. Lie algebra

Proposition 6. The linear space of all vector fields in the region a man-
ifold M is an infinite-dimensional Lie algebra under the commutator of
the vector fields [a, b] = ab - ba.

Proposition 5 implies that the linear space ofall vector fields tangent
to a smooth submanifold is a Lie subalgebra in the Lie algebra of all
vector fields on the manifold.

2.1.5 Visual Material

Figure 2.1.18 presents the graph of a smooth function on the two-


dimensional domain of its definition. One can clearly see the critical
points of the function (maxima, passes, Le. saddles, saddle points). The Saddle points
theory of critical points of functions is one of the essential branches of
modem geometry (see Sect. 2.1.3). The mountain relief in the figure can Mountain relief
be fairly complicated, but if we assume all the interior critical points (i.e. and critical points
those located on the island shore) to be isolated and calculate the number
A of maxima, the number B of minima and the number C of saddles,
then of these numbers we may construct the number A - B + C which
always proves to be equal to unity for an island homeomorphic to a disc.
A mountain relief (or tongues of flame which can always be regarded
as a relief, an interface) of a more complicated structure is depicted.
Figure 2.1.19 shows the structure of a three-dimensional manifold with
boundary in a neighbourhood of a boundary point. Such a point is always
the centre of the equatorial cross-section of a ball, the neighbourhood of
the point consisting of half the ball. In the foreground are shown usual
balls, i.e. the neighbourhoods of the interior points of the manifold. These
points are centres of the balls.

Page 96, Fig. 2.1.18. Morse functions, polynomials of high degree. Ancient
religious cults devoted to underground flame and volcanos eruption; the
origin of monotheism

Page 97, Fig. 2.1.19. Algebraic surfaces and singular points. Ancient tem-
ples and ruins. Greek legends, Persians wars
98 2. Low-Dimensional Manifolds

2.2 Visual Properties of One-Dimensional Manifolds

2.2.1 Isotopies, Frames

We recall the definition of a manifold. A set of points M in a Euclidean


n-manifold space is called an n-dimensional manifold, or simply an n-manifold if
each of its points has a neighbourhood homeomorphic to a Euclidean
space JRn.
In other words, a manifold is a local Euclidean space. The space JRn
is homeomorphic to an open n-ball, and therefore a topological space
is a manifold if each of its points has a ball neighbourhood. To avoid
pathological examples, one customarily imposes tow additional condi-
Hausdorff topological space tions. First, it is required that the manifold be Hausdorff topological
space. This means that any two of its points must have non-intersectiong
neighbourhoods. That is, it is required that any of these points can be
encircled with "its own" ball in such a manner that these balls will not
intersect. For subsets of a Euclidean space (see above) this condition is
automatically fulfilled. Second, it is required that the manifold can be
covered with not less than a countable number of ball neighbourhoods.
In what follows we consider only such manifolds without specifying this
fact specially each time.
One-dimensional manifolds There exists only two connected one-dimensional manifolds - the
circle 8 1 and the straight line JR I • In Fig. 2.2.l it is clearly seen that each
point of the circle or of a straight line has a neighbourhood homeomor-
phic to an open I-ball (Le. to an interval).
As we already know, each smooth manifold, including one-dimen-
sional, can be smoothly embedded (or immersed) into a certain Euclidean
Fig. 2.2.1 space. Embedding a manifold into JRn, we represent it in the form of a
smooth surface in a Euclidean space. Such visualization often proves to
be useful in solving many practical problems. At the same time it is clear
that one and the same manifold can be, generally speaking, embedded
into one and the same Euclidean space in various ways. A natural ques-
Non-equivalent embeddings tion arises: in how many "essentially different" (non-equivalent) fashions
and immersions can one embed a given manifold into a fixed Euclidean space (or more
generally, into another fixed smooth manifold)? The answers to such
questions are given by the manifold embedding and immersion theories
2.2 Visual Properties of One-Dimensional Manifolds 99

developed in modem topology. We cannot present here a systematic pre-


sentation of these interesting and at the same time rather complicated
problems. Therefore we shall only give several visual examples.
We begin with a particularly simple case of one-dimensional mani-
folds. As we already know, according to the Whitney theorem any one-
dimensional manifold can be embedded into a two-dimensional Euclidean
plane. Indeed, it is not at all difficult to construct such embeddings. They
are, in fact, constructed in Fig. 2.2.1. The next question is: in how many
different ways can a circle or a straight line be embedded into a plane?
To answer this question, one should first decide what embeddings are to
be thought of as identical, i.e. equivalent. First we recall the concept of
homotopy.

Definition 1. Let X and Y be topological spaces and let I be a unit


segment. The maps i : X - t Y and g : X - t Y are called homotopic
(f tv g) if there exists a continuous map F of the direct product F : Xxo
X x I - t Y, such that on the bases of the cylinder X x I it coincides Fig. 2.2.2
with the maps i and g. If X and Yare smooth manifolds and i, g, F
are smooth maps, then F is called a smooth homotopy. Given this, the Smooth homotopy
maps i and g are called smoothly homotopic. Two spaces X and Y are
called homotopy equivalent if there exist continuous maps i : X -t Y
and g : Y - t X such that hi is homotopic to the identity self-map X,
i.e. lx, and ih is homotopic to ly (Fig. 2.2.2).

This definition can be reformulated. The maps i and g are homo-


topic providing there exists a family of continuous maps 'Pt : X -t Y
(continuously depending on t and x E X) "joining" the maps i and g,
i.e. if at the initial moment of deformation t = 0 we have 'Po = i and Fig. 2.2.3
at the fmal moment of deformation t = I we have 'PI = g. It suffices to
put 'Pt(x) = F(x, t). See Fig. 2.2.3. All continuous maps homotopic to a Homotopic mappings
given one form a class ofpairwise homotopic maps (or a homotopy class Homotopy class
of a given map). Suppose we are given two embeddings i and g of the
manifold X into the manifold Y.

Definition 2. Two smooth embeddings i, g : X - t Yare called isotopic Isotopic embeddings


(or smoothly isotopic) providing there exists a smooth homotopy F :
X x I - t Y between the maps i and g, such that the map 'Pt : X - t Y
100 2. Low-Dimensional Manifolds

described by the formula F(x, t) = 'Pt(x) is a smooth embedding for each


t E [0, 1].

In other words, two embeddings are isotopic if they can be deformed


Embedding class into each other within the embedding class, i.e. so that in the course of
deformation the manifold 'PtM remains smoothly embedded in N.
Equivalent immersions Similarly, two immersions of the manifold M into the manifold
N are called equivalent if they can be smoothly deformed within the
immersion class, i.e. if there exists a smooth homotopy 'Pt : M - N
such that each 'Pt,O ::; t ::; 1 is an immersion.
Returning to the question formulated at the beginning of this sub-
section, we assume two embeddings of the given smooth manifold M
into the manifold N to be equivalent, "identical", providing these em-
beddings are isotopic. Otherwise, two embeddings are thought of as non-
equivalent, "distinct".
It can be readily seen that any two smooth embeddings of a straight
line into a plane are isotopic. Analogoulsy, any two smooth embeddings
of a circle into a plane are isotopic. Thus, a one-dimensional mani-
fold Ml can be embedded into a plane only in a unique manner (up
to isotopy). This fact can be illustrated as follows. Let f and 9 be two
embeddings of a straight line ~I into a plane ~2. Now consider in ~3
two parallel planes IIo and III and in each of them consider the image of
Fig. 2.2.4 the straight line ~I under the embeddings f and 9 (Fig.2.2A). Consider
a variable plane lIt, where 0 ::; t ::; 1, parallel to the planes IIo and
III which moves from IIo to III with t varying from 0 to 1. Placing
the curve 'PtMI into the plane IIt and varying t, we obtain a cylinder
(a two-dimensional surface) whose base coincides with the two given
embeddings of the manifold MI into the plane.
Non-equivalent At the same time there exist many distinct, non-equivalent immer-
immersions of a circle sions of a circle into a plane. Consider, for example, two immersions
into a plane shown in Fig.2.2.5. It is intuitively clear that remaining within the im-

0 1 00
mersion class one cannot deform a standardly embedded circle into a
figure-of-eight. Any attempt to make such a deformation will necessar-
ily meet with an obstacle, also shown in Fig. 2.2.5. One can see that

9Q QQ2z?
there definitely appears a decreasing loop which fmally degenerates into
a sharp beak. In the vertex of this beak the smooth map of the circle
2.2 Visual Properties of One-Dimensional Manifolds 101

into the plane is not an imm~rsion (verify it by calculation!). Therefore


there are more immersion classes of a circle into a plane than embedding
classes.
We now consider finer properties of embedded manifolds. Let Mk
be a smooth submanifold in a Euclidean space IRn+k. Let at each point x
from Mk a normal (orthogonal) frame T(X) to the submanifold Mk be
given, the frame depending smoothly on the point. The reference frame
T(X) consists of independent vectors e) (x), ... ,en(x) orthogonal to the
submanifold M as shown in Fig. 2.2.6. In this case we shall say that along Fig. 2.2.6
the submanifold Mk an n-frame field T is given to" equip the manifold
Mk. In such a case, the manifold M is sometimes referred to as aframed Framed manifold
manifold and the field T as a frame.
If the submanifold Mk in IRn+k is fixed, it can be equipped with
a frame (i.e. be transformed into a framed submanifold) in many ways.
We shall consider the simplest case of embedding a circle into a plane.
From the intuitive point of view there obviously exist here two distinct
ways of equipping a circle with a frame. To do so, one should direct the
normal field either outside or inside the circle (Fig. 2.2.7). But in order
that we can distinguish between frames, we should first give a defIni- Fig. 2.2.7
tion or identical, or equivalent frames. We shall dwell on the following

@
natural defInition. We shall say that two frames T and'TJ of one and the
same submanifold Mk in IRn+k are equivalent if there exists a smooth 7
deformation of the submanifold Mk in IRn+k within the immersion class ::f.
transforming the normal frame field T into the normal frame field 'TJ. It is
required that the field I.{JtT (which is the result of a smooth deformation
of the initial field T) be normal to the immersed submanifold I.{JtM, i.e.
that I.{JtT be the frame of the manifold I.{JtM for each t varying from 0
to 1.
We shall again examine the two frames of a circle shown in
Fig. 2.2.8, i.e. the normal field outside and the normal field inside. The
question is whether these two frame are equivalent, i.e. whether it is
possible to turn the circle inside out, to make its interior become ex-
terior and vice versa. We can prove these frames to be non-equivalent
in the sense of the definition given above. Intuitively this follows from
Fig. 2.2.8 which shows an attempt to deform the frame T into the frame
'TJ. Pressing two opposite arcs towards each other, we try to get the larger Fig. 2.2.8
102 2. Low-Dimensional Manifolds

part of the nOrm;ll exterior vector field directed inside the curve ob-
tained. It is obvious, however, that in this case there necessarily arise
and remain some small loops on which the field is directed outside as
before. Decreasing in size, these loops have fmally to degenerate into
sharp beaks forbidden by our rules so long as in the vertex of each such
beak a smooth map of the circle is already not an immersion (a break of
the curve).

2.2.2 Visual Material

Figure 2.2.9 presents two human bodies resembling to composition from


the centre of Rembrandt's canvas "The Return of the Prodigal Son". But
this image is far from being immediately guessed since the whole of the
composition was subjected to a homeomorphism making the original pic-
ture unrecognizable. This is the important property of homeomorphism
Metrical properties - to change the metrical (accordingly, easily recognizable) properties of
the object and to retain its topological (not so easily recognizable) prop-
erties. In the background, behind the horizon, there are framed circles.
In Fig. 2.2.10 we can see a human face with details reproducible in any
photo. At the same time it is known that textual and visual informa-
tion is often excessive. For example, from a literary connected text one
can remove fragments of many words and sometimes whole words from
word combinations and set expressions without loss of the meaning of
the text. In exactly the same manner one can remove many fragments
Recognition problem from a visual image but the object will remain recognizable. It is, of
course, desirable to remove only insignificant details and to retain the
essence. This is just what homeomorphisms and homotopies do. They
make it possible to recognize the most important thing in a geometrical
object, to free it of the mass of secondary unimportant details which of-
Invariants of the object ten hamper recognition of the principal invariant properties of the object.
Figure 2.2.11 presents the female portrait deprived of some habitual de-
tails but retaining nonetheless the most essential characteristic features of
Visual perception the original. This shows how resistant the visual perception is to ''white
noise". This is explained by profound properties of perception first of
all evidently fixing invariant principal properties of the object and then
supplementing them with some minor details of metric character. This
~

Page 103, Fig. 2.2.9. Homeomorphism preserves the topological properties


of an objects framed submanifolds; Mathematical variations on the idea of
Rembrandt's 'The Retum of the Prodigal Son" .
104 2. Low-Dimensional Manifolds

idea provides a deeper insight into the mathematical nature of homeo-


morphism and homotopy equivalence. Another "portrait" whose features
were subjected to a still stronger "erasure" is shown in Fig.2.2.12. It
should be emphasized that this significant fact of recognizing an object
already from some details underlies various modem statistical pattern
Recognition methods recognition methods. The basic problem in the development of corre-
sponding methods is recognition of principal invariant characteristics of
the object. The solution of this problem is most often nontrivial. We
have tried to illustrate the same idea in Fig.2.2.13 using mathematical
fantasia based on the analysis of some paleontglogical data. It would not
Reconstruction be out of place to recall here the science of reconstruction of the image
of the image of ancient organisms (which do not exist now) developed by modem
paleontology. Experts have to reconstruct the image of a creature having
at their disposal not more than fragments of a skeleton or even bones of
animals accidentally preserved to these days. See Fig. 2.2.14. Motivation
of pattern recognition from partial information is widely used in modem
arts, including ballet, for example, so-called "Sculpture Theatre". See
Fig. 2.2.15.

Page 105, Fig. 2.2.10. Homeomorphism and homotopy of geometrical


objects; recognition of topological invariants. German and Scandinavian
myths on Alberich (legend about Siegfried)

Page 106, Fig. 2.2.11. Recognition of the prinCipal invariant properties of


the object. The fantastic female portrait (Elena Kuz'menko)

Page 107, Fig. 2.2.12. Topological deformation and invariants of a homeo-


morphism. Ancient Australian myths

Page 108, Fig. 2.2.13. Reconstruction of an object using statistical pattern


recognition methods. Ancient Roman legend about Romulus and Remus;
foundation of the Rome

Page 109, Fig.2.2.14. Homotopy equivalence and reconstruction of the


images based on geometrical invariants. Medieval Russian legends; Gog
and Magog, Rush and Mesheh

Page 110, Fig. 2.2.15. Idea of a smooth isotopy and singular points of
algebraic surfaces. Stone monument-temple; pre-Christian cult
2.3 Visual Properties of Two-Dimensional Manifolds 111

2.3 Visual Properties of Two-Dimensional Manifolds

2.3.1 Two-Dimensional Manifolds with Boundary

As we have already seen, there are few I-manifolds: all of them are
unions of circles and straight lines. There are much more 2-manifolds:
for example, any open subset of a plane is a 2-manifold (Fig. 2.3.1). We
shall give examples of compact 2-manifolds: the sphere 8 2, the torus
T2 = 8 1 x8 1 (Fig. 2.3.2). The sphere gn and the torus rn =T n - I x8 1 are Fig. 2.3.1
examples of compact n-manifolds. Other examples of multidimensional
manifolds can be constructed noting that the direct product of manifolds Direct product
of dimensions m and n is a manifold of dimension m + n. As we have of manifolds
seen, it is very convenient to extend the general concept of a manifold
including here the concept of a manifold with boundary. Recall that the
Hausdorff topological space is called an n-manifold with boundary if
each of its points has a neighbourhood h.omeomorphic either to a space
]Rn (i.e. a ball neighbourhood) or to a closed half space ~ (i.e. a half-
ball neighbourhood). The union of points of a manifold M which have
no neighbourhood of the first type is called its boundary 8M. Fig. 2.3.2

Theorem 1. If the boundary of an n-manifold is non-empty, it is an Boundary


(n - I)-manifold without boundary. of the manifold

The proof of this theorem is quite obvious since all the points of
the boundary hyperplane of the space ~ are equivalent and have in
this hyperplane (n - I)-ball neighbourhoods. On the other hand, the
intuitively obvious existence of at least one manifold with a non-empty
boundary is a rather nontrivial fact from the formal logical point of view
since it is not a priori clear why the half space ~ (with a distinguished
point on the boundary) is not homeomorphic to the space ]Rn (with a
distinguished point inside). This is, in fact, a direct consequence of the
Brouwer theorem on invariance of a region (see Ref. [36], vol. I). In line Brower theorem
with our principle we do not dwell here on thorough investigations to
confirm the intuitively clear (and valid) facts. One should however bear in
mind that there exist seemingly obvious but of course incorrect (I) mathe-

.0' : .,
.. <:
:::0"
...--... <:.... .::~.;:
112 2. Low-Dimensional Manifolds

matical facts. We shall each time infonn the reader which "geometrically
obvious" facts are valid and which are not.
We shall give examples of manifolds with boundary: the half in-

Q2) :.:;. :. ":':::":' Ir.=-.t. terval (a noncompact manifold) and the segment (a compact manifold).
0:':::::;"
'. \'::::::": . :;:" :;;
These are one-dimenional manifolds. Next: the disc D2, the ring 8' x I,
and the Mobius strip (two-dimensional manifolds), see Fig. 2.3.3. Their
. ; :";~": .::":: . ~::. boundaries consist of a point, two points, a circle, two circles and a circle,
respectively.
Fig. 2.3.3
We consider a three-dimensional ball, i.e. a set of points bounded
in 1R3 by a standard sphere. The direct product of a disc by a segment,
i.e. a filled cylinder, will be called a three-dimensional handle. Clearly,
such a handle can be glued to a ball along the two bases of the handle,
i.e. along two discs. As a result, we obtain a ball with a handle. One can
glue several such handles to obtain a three-dimensional body called a full
Solid pretzel/solid) pretzel or a filled (solid) pretzel, the number of handles being the
genus of the pretzel. Thus, an example of a 3-manifold with boundary
may be a full pretzel Kp of genus p (a ball with p handles as shown in
Fig. 2.3.4). For any n, and n-dimensional ball Dn and a half space ~
are examples of n-manifolds with boundary.
The direct product of manifolds of dimensions m and n with bound-
ary is a manifold of dimension m + n with boundary. Given this, there
holds the fonnula
a(M x N) = aM x N U M x aN

resembling the fonnula of differentiation of a product of functions. Note


Fig.2.3.4 that this analogy is not limited to the outward resemblance.
Let M, and M2 be n-manifolds with homeomorphic boundaries and
let h: aM, - aM2 be a homeomorphism. If we glue the manifolds M,
and M2 by the homeomorphism h, i.e. identify each point x E aMI with
the point h(x) E aM2, we obtain an n-manifold M without boundary.
Indeed, each point x E aMI in M, has a half-ball neighbourhood, the
point h(x) has a half-ball neighbourhood in M 2. These half-ball neigh-
bourhoods are glued together to give the ball neighbourhood of the corre-
sponding point in M (Fig. 2.3.5). For example, if we remove an open disc
from a sphere and glue in a Mobius strip through a certain homeomor-
phism of their boundaries (circles), we obtain a new 2-manifold called
Fig. 2.3.5 a projective plane and denoted by RP2 (Fig. 2.3.6). The projective plane
2.3 Visual Properties of Two-Dimensional Manifolds 113

is not embedded into 1.3. One can also glue together n-manifolds along
submanifolds of their boundaries to obtain n-manifolds with boundary.
.. .
@
' ... .
For example, a solid pretzel K2 of genus 2 can be obtained by glueing to- .. . .
" ',

gether two solid pretzels of genus 1 (full tori) through a homeomorphism


of discs on their boundaries (Fig. 2.3.6).

2.3.2 Examples of Two-Dimensional Manifolds

The simplest closed 2-manifold is a sphere. Besides, we are already


acquainted with a torus which admits also a representation other than SI x
SI. We have already mentioned that the process of constructing mani-
folds resembles to some extent the preparation of papier-mache when a
clay dummy is several times covered with small pieces of paper glued F'Ig...
236
on it. Since these pieces are very small, one can using them reproduce
rather complicated clay models. Bearing in mind this analogy, we shall
presently restrict ourselves for simplicity to a one square sheet of paper.
The question is what two-dimensional manifolds can be obtained from
this sheet by glueing the boundary edges, the sides of the square. For
convenience of the description, we fix (say, clockwise) orientation on
the square, assign to each edge on the boundary a letter and provide the
edge with orientation (drawn an arrow). This orientation may coincide
or not coincide with the direction of motion induced on the edge by the
orientation of the square. We choose an arbitrary vertex of the square
(e.g. the left lower one) and moving clockwise will successively write
out the encountered letters aT in powers e = =Fl, indicating by the power
e = +1 the fact of coincidence of the arrow on the edge with the direction
of our motion and by the power e = -1 the fact of the opposite direction
of the arrow. As a result, we obtain a word W composed of letters. See
examples in Fig. 2.3.7. The corresponding words W have the form: 1)
abb-1a- l , 2) aba-1b- l , 3) abab, 4) abab- I • What manifolds have we
obtained? In Fig. 2.3.8 one can see that the glueing of a square according
to the requirements indicated in the notation of the word W results in the
ordinary sphere S2. For convenience we have replace the initial square
by a homeomorphic object, namely, a sphere with a rectangular hole cut
out, on whose sides the same letters and arrows are placed (Fig. 2.3.8).
114 2. Low-Dimensional Manifolds

We shall not describe the details of the glueing in words, but suggest the

Cl[J:
.. ·\::l .CD·}::·: a
well-known formula used by medieval geometricians: Vide! (see).
The second word in Fig. 2.3.7 determines the glueing process visu-
. . .... ." .
••• • '0 •••••••••

. . ally shown in Fig. 2.3.9. We have obtained a two-dimensional torus (a


a $1 -8 r:L roll). Thus, in the fIrst two cases we have got two-dimensional orientable
manifolds smoothly embedded in 1.3. Their orientability readily follows

·ciJ· ·EJ·
from the defInition (we leave this exercise to the reader).
The third word in Fig.2.3.7 determines a projective plane JlU>2. To
make sure of this, we recall the classical defIntion of a projective plane.
Its points are various straight lines l in R3 passing through a fIxed point,
:g RP2. -8 K2- for example, the origin (Fig. 2.3.10). An equivalent model is constructed
Fig. 2.3.7 like this. We consider a standard sphere centred at the origin. Then
each straight line from our sheaf is uniquely defIned by points of its
intersection with the sphere. There are exactly two such points. They are
diametrically opposite on the sphere. We shall denote them by (x, -x).
So, a projective plane can be given as a set of pairs of the form (x, -x),
where x runs through the sphere. An equivalent model is a projective
plane obtained from a sphere 8 2 through identifIcation of its diametrically
opposite p~ints. This can be conditionally written as RP2 = 821'l2. The
point is that on a sphere one can defme a smooth action of the group 'l2
consisting of two elements: unity and involution (J, i.e. the transformation
whose square is equal to unity (identity transformation). The involution
is given by the formula (J(x) = -x. Factorizing the sphere by the action

:tillj
.
of this group (i.e. identifying the points x and (J(x», we just obtain a
projective plane.
-8 •.•.•. The next equivalent model of a projective plane is this. Let us con-
~.:«.~. -- -
t1 ::
sider the upper (or lower) hemisphere 81 and identify on its boundary,
i.e. on the boundary circle, diametrically opposite points (Fig. 2.3.10).
: :: '-8'. :-....... :
'.
".
.
'0··
II
b' ::
::
Equivalence of this model to the preceding ones is obvious. It suffIces
..... ·f· ~ to cut the sphere by a plane passing through the centre of the sphere.
Discarding the lower hemisphere, we fInd out that to construct RP2 it
suffIces to take only the upper hemisphere, but one should identify dia-
metrically opposite points on the equator.
This model can be reformulated. Applying homeomorphism (assum-
ing the sphere to be made of thin rubber which can be expanded without
ruptures and glueings), we deform the hemisphere into a large sphere
Fig. 2.3.9 from which a small disc is discarded, and then on the boundary of the
2.3 Visual Properties of Two-Dimensional Manifolds 115

hole obtained we identify diametrically opposite points (Fig. 2.3.10). Fi-


nally, we describe the last model. The same hemisphere can be deformed
homeomorphically into a flat two-dimensional disc on whose boundary
the opposite points are identified (Fig. 2.3.11). A disc is homeomorphic
to a flat square. To describe the identification on the boundary of the
square of points symmetric relative to the centre of the square, it suf-
fices to place letters a and b on the sides, as shown in Fig. 2.3.11. As a
result, we obtain a word abab indicated above as one of the codes of a
projective plane.
The question of modelling (more precisely, smooth immersion) of
a projective plane in 1 3 is discussed below. We now proceed to the
last, fourth code-word in Fig. 2.3.7. The surface obtained from a square
through its glueing accordings to this code is called a Klein bottle K2 and
(as a projective plane) is a non-orientable closed 2-manifold. Construction
of K 2· is analogous to glueing a torus from a square, but at the end of the
construction we meet with an important difference: we now cannot glue
the bases of the cylinder in 1 3 without self-intersection of the surface. In
Fig. 2.3.12 one clearly sees the precise manner in which the last glueing
should be made. Thus, we have succeeded in modelling a Klein bottle
in the form of a closed surface in 1 3• It has a self-intersection, but at
the same time it is clear that we have constructed an immersion of K2
into 13.
How shall we prove that neither the projective plane nor the Klein
bottle can be smoothly embedded into 13? It suffices to make sure that
each closed two-dimensional surface smoothly embedded in 1 3 will nec- Fig. 2.3.11
essarily split ~3 into the interior and exterior parts each of which is the
three-dimensional manifold with boundary. Each of these manifolds is
orientable, and since the events occur in 1 3, it follows that on their
boundary the concept of the inward and outward normals to the surface
is uniquely defmed. This implies that the initial two-dimensional surface
is orientable.

2.3.3 Modelling of a Projective Plane in a Three-Dimensional Space

We now return to immersion of a projective plane into 1 3• First we shall


try to model "at least roughly" tp2 in the form of a surface in 1 3. We Fig. 2.3.12
116 2. Low-Dimensional Manifolds

take a square corresponding to code-word N° 3 in Fig. 2.3.7, i.e. abab.


Assuming the square to be made of thin rubber, we press it downwards
behind the plane of the book sheet and obtain a sphere from which a
small square with the indicated glueings on the sides is cut out. Further
events are shown in Fig. 2.3.13 ("Vide!"). As a result, we obtain a rather
complicated surface with self-intersection. For obviousness, the figure
shows its cross-sections by planes orthogonal to the singular segment
in the upper portion of the figure (and, thus, to the plane of the book).
Clearly, the constructed map f of a projective plane in R3 is not even an
immersion, to say nothing of an embedding. The obstacle is the ends of
the singular segment AB on which two sheets of the surface intersect, the
intersection being such that at the points A and B there arises a complex
structure distinct from the intersection of several locally Euclidean discs.
Nonetheless, the model constructed has made it possible to "have a look"
at some (although not very "good") picture of a projective plane in a
three-dimensional space. Figure 2.3.14 presents evolution of the cross-
sections of the figure constructed and its penultimate stage (N° 4 in
Fig. 2.3.13
Fig. 2.3.13) by planes orthogonal to the singular edge AB. This surface
can be giv~n in R3 by the algebraic equation (ax 2 + by2)( x2 + y2 + z2) =
2z(x2 + y2).
We shall not, of course, stop at this and shall now construct an
immersion of][Ul2 into R3, We shall preliminarily describe several simple
facts from visual topology which are instructive for what follows.
Mobius strip We shall consider the Mobius strip J.L shown in Fig. 2.3.15. It is
obtained by glueing opposite sides of the square with orientation reversal.
It turns out that when glueing the Mobius strip wich a disc, we obtain a
projective plane.

Lemma 1. A projective plane is homeomorphic to the manifold obtained


by glueing a Mobius strip with an ordinary disc along their common
boundary. In other words, discarding a two-dimensional disc from ][Ul2,
Fig. 2.3.14
we get a Mobius strip. This fact will be conditionally written as ][Ul2 =
J.t+D2,

Proof. Figure 2.3 .16 presents a model of a projective plane in the form
of a square with glueings abab. We shall prove that when discarding a
Fig. 2.3.15 disc from [U>2, we obtain a Mobius strip. Since lR.p2 is a smooth closed
2.3 Visual Properties of Two-Dimensional Manifolds 117

manifold, it does not make any difference where we choose the centre of
the discarded disc. At this moment all the points of the projective plane
are equivalent. We choose the point in the middle of the side a. The
disc centred at this point will be depicted as its two half discs (black in
the figure) glued together along the diameter. Discarding this disc, we
obtain, after straightening the remaining angles, a square whose opposite
sides are glued with orientation reserval, as required.

In the classical (usual) embedding of a Mobius strip into R3, its


boundary, i.e. a circle, is so embedded that it goes twice around the
vertical axis (Fig. 2.3.l7). We shall try to deform the Mobius strip in R3
so as to simplify the embedding of its boundary. Namely, we shall make
the boundary of the Mobius strip depicted by an ordinary flat circle, i.e. Fig.2.3.16
a circle located in a plane. This can be done only through complication
of representation of the Mobius strip itself. How will it be positioned?
The answer appears to be given by Fig. 2.3.17. The pattern presented
is obviously obtained from the model of the projective plane shown in
Fig. 2.3.13 by singling out its lower cup by a plane orthogonal to the
plane of the book sheet. Cutting out the cup we, in fact, discard a disc
from the projective plane. The remaining part (with self-intersection) is
occasionally referred to as a crossed cap. So, a crossed cap is simply a
Mobius strip so located in R3 that its boundary is a standard flat circle.
Fig.2.3.17
Lemma 2. A Klein bottle is obtained by glueing together two Mobius
strips along their boundary. In other words, cutting a Klein bottle along
an appropriate circle we have two Mobius strips. This fact is condition-
ally written out as follows: K2 = JL + JL.

Proof. We take a square with glueings abab- 1, which represents a Klein


bottle. In the middle of the side a we mark a point 1/2 and join it with
the vertices of the square as shown in Fig. 2.3.18. Then we cut the square
along these two lines. As a result, we obtain a parallelogram and two
triangles. The parallelogram is obviously a Mobius strip. Glueing the
two triangles along their common side b, we obviously obtain a second
Mobius strip. This proves the lemma.
118 2. Low-Dimensional Manifolds

How shall we depict this line of the cut on a Klein bottle immersed
in 1R.3 (Fig. 2.3.12)? We invite the reader to trace out attentively the cut
line motion in Figs.2.3.18 and 12. So, cutting a Klein bottle as shown
in Fig. 2.3 .19, we get two Mobius strips. Each of them is immersed in
lR3 (with self-intersection). Consequently, this model of a Mobius strip
is more regular (a ''better'' one) than a crossed cap which, we recall,
is not an immersion of the Mobius strip. Note that the immersion of
the Mobius strip discovered by us is remarkable for the fact that its
boundary is immersed into a plane, i.e. its boundary is a flat curve (with
self-intersections).
Fig. 2.3.19 According to Lemma 1, a projective plane is obtained by glueing
a Mobius strip with a two-dimensional disc. By proving Lemma 2, we
have constructed a Mobius strip immersion such that its boundary is a flat
circle. We shall use this immersion to continue it to the immersion of the
entire projective plane. We assume this Mobius strip immersed in lR3 to
be part ofprojective plane. To immerse the whole of the projective plane,
Fig. 2.3.20 it suffices to add a disc, i.e. to glue the flat boundary of the Mobius strip
by the immersed disc. So, we are approaching the following problem.
We are given a flat curve 'Yo shown in Fig. 2.3.20 which is an immersion
of a circle into a plane. How shall we glue it by an immersed disc?
Let lRij be a plane containing the Mobius strip boundary (Fig. 2.3.21).
We start moving this plane raising it up parallel to itself. This fives a
family of planes ~~ depending on the parameter t, where 0 ~ t ~ 1.
Simultaneously we start smoothly deform the initial flat curve 'Yo in this
plane. This gives a family of curves 'Yt. With varying t, these curves form
a two-dimensional surface (sweep up the surface). We shall describe the
Fig. 2.3.21 process of deformation of the curve 'Yo into curves 'Yt. See the illustration
of this process in Fig. 2.3.22. This smooth deformation leads to the fact
that the curve is slipping down itself and becomes a curve already without
self-intersection, i.e. a circle embedded standarly into the plane. As has
been mentioned above, as the plane goes up, this deforming curve sweeps
up a certain surface. At the moment when the curve becomes a circle we
glue it by a smooth disc. Thus, we have glued the Mobius strip with a
disc, i.e. obtained a model of a projective plane. At the same time we have
Fig. 2.3.22 obviously constructed its immersion into lR3• How is the set of points of
2.3 Visual Properties of Two-Dimensional Manifolds 119

this immersion organized? It follows from Fig. 2.3.22 that the set consists
of three circles glued at one point. As a result, we obtain a surface in
space representing an immersed projective plane. This realization of a
projective plane is called the Boy surface [38]. It can be drawn in JR.3. To
do so, we represent the projective plane as a result of identification of
opposite points on the boundary of a regular hexagon (but not a square, as
we have done before). See Fig. 2.3.23. The corresponding code is abcabc.
Clearly, a hexagon with such identifications is equivalent to a disc with
opposite points of the boundary identified. As before, we transform the
hexagon into a sphere from which a hexagon is cut out (Fig. 2.3.23).
On the sides of the hexagon there stand letters and arrows determining
the glueings. We cut the sphere into three lobes with meridians q, d,p
(Fig.2.3.23). As a result, the sphere falls into three congruent pieces
which coincide with one another under an appropriate turn. We take one Fig. 2.3.23
of these pieces, for example, the one between the meridians q and d
(Fig. 2.3.24). There appear three ascending peaks. We glue them at one
point and denote the point by N (the analogue of the north pole). Note
that this glueing is not needed from th~ viewpoint of the topology of
a projective plane, but as we shall see below, we shall have to glue
the indicated points since we attempt to realize a projective plane in a
three-dimensional space. The result is shown in Fig. 2.3.24, Step 2.
We now fix two meridians q and d and raise the loop a as shown
in Fig.2.3.24, Step 3. The surface will stretch to become a dome- or
shell-shaped.
The next step is to tum the loop c to the right top side taking it
inside the dome and nearer the inner side of the dome in such a way that
the loops a and c occupy the position indicated in Fig.2.3.24, Step 5.
Step 6 in the same figure is the final shape of the surface. Here the arcs
d and q are congruent, and when the whole figure turns through an angle
of 27r /3 around the vertical axis SN, the arc d is superposed on the arc
q and the loop a on the loop c. Similar deformations will be performed
for the other two lobes of the initial surface. As a result, we obtain three

3
120 2. Low-Dimensional Manifolds

congruent figures. We take the second lobe. The corresponding letters


on it will be primed. We make part of the boundary of the first lobe
coincide with the secone one. To this end we match the arcs d' and q so
that the point 8' meets the point 8 and the point N' meets the point N.
Consequently, the loop a' coincides with the loop c. What is the boundary
of the surface obtained, i.e. of the glueing together of two lobes? Clearly,
the boundary consists of the arcs q', a, c', d. To make this process more
C demonstrative, we go back to the initial hexagon on which we mark all
Fig. 2.3.25 the cuts participating in the glueing (Fig. 2.3.25). Three hexagon edges
emanating from its centre are endowed with a pair ofletters indicating the
edges glued under restoration of the projective plane of the three lobes.
Step 6 in Fig. 2.3.25 shows the vertical dashed third-order symmetry axis.
This means that when the figure turns round this axis through an angle
of 27l' /3, the edge d coinsides with q and the loop a with the loop c. So,
to the two globes already glued we add the third lobe according to the
glueings determined by the hexagon in Fig. 2.3.25. As a result, we obtain
a Boy surface (Fig.2.3.26). It is clearly seen that we have constructed
an immersed surface, i.e. as a point of the surface varies, the tangent
plane at this point varies continuously (and smoothly). The surface has
no breaks or other singular points. It has only several self-intersection
lines in the form of three loops. Two sheets of the surface intersect in
the neighbourhood of each point on these loops, different from the point
Fig. 2.3.26 N (the centre of the figure).

2.3.4 Two Series of Two-Dimensional Closed Manifolds

In the preceding subsections we have discussed the properties of the


surface obtained by glueing the sides of a square. We now consider a
more complicated flat polygon, for instance, an octagon on whose sides
Coding of we place letters and arrows according to the code aba-1b-1cdc1d- l .
two-dimensional surfaces From the algebraic point of view, we have multiplied to commutators
[a, b] = aba-I b-I and [c, d] = cdc 1d- I . Recall that the commutator
[a, b] determines a torus (see above). This analogy with algebraic objects
is not exhausted by an outward resemblance but is a consequence of deep
Fundamental group relations between the code word Wand the fundamental group of the
2.3 Visual Properties ofTwo-Dimensional Manifolds 121

Fig. 2.3.27

manifold. We shall perform all the glueings required by the code word
W (Fig.2.3.27). The manifold obtained is called a pretzel of genus 2
or simply a pretzel. It admits another representation which we present
below. First recall the operation of glueing a handle. A handle is a usual
cylinder (Fig.2.3.28). Its boundary consists of two circles. We discard
two non-intersecting discs from a two-dimensional manifold to obtain a
manifold with boundary {two circles}. We glue a cylinder to this manifold
by identifying its boundary circles with the boundaries of the holes in
the manifold. This operation, resulting in a new manifold, will be called
the glueing of the handle.

Lemma 3. A torus is obtained from a sphere by means of glueing one Fig. 2.3.28
handle. A pretzel is homeomorphic to a sphere with two handles.

The proof is shown in Fig. 2.3.29. The notation is glueing a handle


will be as follows: a torus = T2 = S2+r; a pretzel = S2+2r, where r is a
handle. It is clear that we now can construct an infinite series of manifolds
Fig. 2.3.29
122 2. Low-Dimensional Manifolds

by glueing an arbitrary number of handles to a sphere (Fig. 2.3.30). We


obtain the first series of two-dimensional Mi = 8 2 +gr. A non-negative
integer 9 is called the genus of the swface. These manifolds can be
represented as a result of glueing polygons by analogy with what has
been done above for a torus and a pretzel. Consider a 4g-gon on a plane.
On the sides of this rectangle there are letters arranged according to the
code

Fig. 2.3.30 i.e. we multiply 9 commutators.

Lemma 4. The manifold obtained by glueing the sides of a 4g-gon ac-


Code of the sphere cording to the indicated code Wg is homeomorphic to a sphere with 9
with g handles handles.

The number of commutators in the code word Wg coinsides with


the genus of the surface. We leave to the reader the simple verification
of the statement: all manifolds Mi are orientable and closed.
We shall point out the second. infmite series of 2-manifolds. First
we reformulate Lemmas 1 and 2. We define the operation of glueing a
Mobius strip. We discard a disc from the manifold, then glue a Mobius
strip through a homeomorphism of its boundary and the boundary of the
hole obtained (Fig. 2.3.31).

Lemma 5. A projective plane is homeomorphic to a sphere with one


Mobius strip. A Klein bottle is homeomorphic to a sphere with two
Mobius strip.

The proof follows immediately from Lemmas 1 and 2 and from the
defmition of the operation of glueing a Mobius strip. The glueing of k
Mobius strips j.L to a sphere will be denoted by 82 +kj.L, where k > O. All
these manifolds are non-orientable. Indeed, we shall use the defmition
of orientability in terms of orientation of a reference frame transported
Fig. 2.3.31 along a closed path on a manifold. On a manifold Sl +kj.L, where k > 0,
2.3 Visual Properties of Two-Dimensional Manifolds 123

we take a closed path 'Y tmversing exactly one Mobius strip (Fig. 2.3.32).
We state that when a tangent frame moves along a path and then returns
to the initial point, its orientation reverses. This fact is illustrated in
Fig. 2.3.32.
Thus, we have discovered two infmite series of 2-manifolds: 8 2 +gr
(spheres with handles, 9 ~ 0), 82 + kp, (spheres with Mobius strips,
k > 0). Manifolds of the first series are orientable, while those of the
second series are not. Manifolds from distinct series are therefore not
homeomorphic since orientability is an invariant of homeomorphism.
There arises a natural question of why we have not yet mentioned
here a mixed series of manifolds when both handles and Mobius strips
are glued to a sphere. It turns out that such a mixed series does not q
provide anything new: it yields manifolds of the second series.

Lemma 6. A manifold of the form 82 +gr + kp" i.e. one obtained from
a sphere via glueing 9 handles and k Mobius strips (where k > 0), is
homeomorphic to a manifold 82 +(2g +k)p" i.e. to one of the manifolds
from the second, non-orientable series. Fig. 2.3.32

Proof. Consider a Klein bottle and deform it in a three-dimensional space


as shown in Fig. 2.3.33. Thus, it can be represented as a sphere with a
glued handle, but this glueing does not correspond to the d~finition of
Fig. 2.3.28. This one is screwed, i.e. we have glued a screwed handle. In
such an operation one base of the handle is glued to the sphere from the
outside and the other from the inside. As a result, we obtain a surface with
self-intersection since in a three-dimensional space it is impossible to get
inside from the outside without crossing the sphere. This is the result of
three dimensions. If we worked in a four-dimensional space, we could
realize this glueing without self-intersection of the surface. In particular,
a Klein bottle can be smoothly embedded (i.e. without self-intersections) Fig. 2.3.33
into ]R4. This is, however, guarenteed by the Whitney theorem. Thus, a
Klein bottle is homeomorphic to a sphere with a screwed handle. On the
other hand (see above), a Klein bottle is homeomorphic to a sphere with Klein bottle
two Mobius strips. Comparing these facts, we arrive at the assertion: the
glueing to a surface of one screwed handle is equivalent to the glueing
to a surface of two Mobius strip.
124 2. Low-Dimensional Manifolds

We now consider a sphere to which both handles and Mobius strips


are glued. For simplicity we restrict our consideration to the case of one
handle and one Mobius strip (Fig.2.3.34). On a surface we consider a
closed path , starting on one of the bases of the handle then passing
along the axis of the Mobius strip and back to the starting point. We
shall realize the following deformation (homeomorphism) of the surface.
We fix one base (bottom) of the handle and then start moving the
other base (bottom) along the path ,. From Fig. 2.3.24 it is seen that the
handle will start deforming if one of its bases comes up to the Mobius
strip, comes onto it, passes along its axis and leaving the Mobius strip
turns over, i.e. the handle is screwed (I). The further motion of the base
along the path, will return it to the initial place. But now the handle has
turned over. This process can be demonstrated on another more obvious
graph (Fig. 2.3.35). Clearly, when the foot (base) of the handle is moving
along the Mobius strip, the direction of the normal to the Mobius strip is
reversed, which causes handle screwing. But since a screwed handle is
equivalent to two Mobius strips, we can now replace the handle and the
Mobius strip by three Mobius strips (without altering the surface). Hence,
Fig. 2.3.34
in the presence of at least one Mobius strip each handle is transformed
into two Mobius strips, from which Lemma 6 obviously follows.
It has been shown above that all orientable closed 2-manifolds (i.e.
the series S2+gr) are smoothly embedded into 1 3. We shall prove that all
manifolds of the non-orientable series 8 2 +kj.L can be smoothly immersed
(but not embedded!) into 13. We have proved that ]Rp2 = 8 2 + j.L, and
therefore any manifold of the form 8 2 + kj.L is representable in the form
shown in Fig. 2.3.36. As we can see, the manifold falls into the sum of a
2.3 Visual Properties of Two-Dimensional Manifolds 125

sphere with k holes and k copies of a projective plane with a discarded


disc. Each of these summands admits an inlmersion into ]R3. The sphere
with k holes is obviously embedded into ]R3 and the projective plane
with a hole is immersed into ]R3 (see above). Consequently, the whole
manifold 82 + kp, is immersed into ]R3.
The two above-mensioned series of surfaces exhaust the list of all
two-dimensional compact closed connected manifolds. Moreover, the
manifolds 8 2 +gr and 82 +sr are not homeomorphic in case 9 =f sand
the manifolds 8 2 + kp, and 8 2 +lp, are not homeomorphic in case k =f l.
In conclusion we formulate several visual problems. Let us consider
the axis of a Mobius strip (the dashed line in Fig. 2.3.37). We cut it along
this circle. What manifold (orientable or non-orientable) shall we obtain?
(We shall obtain an orientable strip screwed by 27r). We shall consider
the screwed strip from the preceding problem and cut it along the axis.
Prove that we obtain two strips (fmd the angle of their screwing) hooked
in]R3 as shown schematically in Fig. 2.3.37. We continue this process by
cutting each of these strips along the axis, etc. We obtain a set of new
strips. Prove that each strip from this set is hooked by another one. Fig. 2.3.37

2.3.5 Classification of Closed 2-Manifolds Classification theorem


for 2-surfaces
From what have been said above we can derive two series of closed (=
compact without boundary) 2-manifolds: surfaces (boundaries) of pret-
zels Kg, 9 ~ 0, and spheres with k ~ 1 discarded open discs instead of
which k Mobius strips are glued. It turns out that there exist no other
closed connected 2-manifolds.

Theorem 2. Any connected closed 2-manifold is homeomorphic either


to the surface of a full cracknel of some genus, i.e. to a sphere with
handles, or to a sphere with several Mobius strips.
Fig. 2.3.38
We now present the scheme of the proof.
a) Any 2-manifold can be triangulated, i.e. divided into curvilinear Triangulation
triangles so that any two triangles either have no common points or of a surface
intersect at one vertex or along one edge (Fig. 2.3.38). This fact seems
intuitively obvious (and it is valid).
126 2. Low-Dimensional Manifolds

Indeed, a manifold is covered with Euclidean neighbourhoods each


of which, of course, admits triangulation. Thus, the problem reduces to
making the triangulations of the neighbourhoods compatible. In the case
of dimension 2 this can be done using the Jordan theorem or the prop-
Fig. 2.3.39
erties of geodesics (if on the manifold we preliminarily introduce some
Riemannian metric, which is always possible). Figure 2.3.39 presents the
minimal triangulation of a torus T2 represented as a rectangle with op-
posite sides identified. The division of the torus into triangles as shown
in Fig. 2.3.40 is not a triangulation since some pairs of triangles (e.g.
triangles A and B) intersect at two vertices.
b) Instead of each vertex of triangulation we take a disc and instead
Fig. 2.3.40
of each edge a strip. See Fig.2.3.41 where D stands for discs, L for
strips and Z for the remaining parts which is convenient to call patches.
Note that topologically each strip is a rectangle glued to the union of
discs along two opposite sides and each patch is a ring that glues the
circle on the boundary of the union of discs and strips. Hence, each
closed 2-manifold can be obtained as follows: we take k discs, glue m
Fig. 2.3.41 strips to them and glue a patch on each of the n circles obtained on the
boundary of the 2-manifold. In this case we say that the 2-manifold has
type (k, m, n).
c) We prove that each closed 2-manifold has type (1, m, 1). To do
so, we should learn to decrease the number of discs and patches. This can
be easily done. If the number of discs is greater then 1, there exist two
discs D1 and D2 joined by a strip L. Then their union will be assumed
to be a new disc. As a result, the number of discs and strips decreases
and the number of patches remains the same (Fig. 2.3 .42). The number of
Fig. 2.3.42
patches can be decreased using dual consideration: two patches with one
of the strips adjacent to them with this free lateral sides can be united
with this strip and thought of as a new patch.
d) Suppose a closed 2-manifold has type (1, m, 1), i.e. is obtained
from a disc D by glueing m strips and one patch. A rectangle can be
glued to the disc D in two ways: either so that its unidirectional lateral
sides prescribe one and the same direction on the circle aD or so that
they prescribe different directions on the circle aD. The strip obtained
Fig. 2.3.43
will called non-orientable or orientable, respectively. Strips 1 and 2 in
2.3 Visual Properties of Two-Dimensional Manifolds 127

Fig.2.3.43 are orientable and strips 3 and 4 are not. Suppose, all the
strips are orientable. Let L1 be one of them. Then there necessarily
exists a strip L2 joining the two arcs into which the circle aD falls by
the base of the strip L1 (Fig. 2.3.44) since otherwise at least two patches
would be needed. Note that the base of each subsequent strip can be
isotopically moved along the boundary of the union of the disc D with the
Fig. 2.3.44
previous strips (Fig. 2.3.45). The manifold remains unaltered. Using this
operation, one can free from the strip bases the region distinguished in
Fig. 2.3.44 thus separating the so-called pair of crossed strips. Repeating
these arguments we see that it is only in the case of orientable strips
that the manifold is obtained from the disc via glueing several (say, k)
individual pairs of crossed strips and one patch (Fig. 2.3.46). It can be
easily verified that the manifold is then homeomorphic to the pretzel ... . '. . .., ... . .
.......
" " '
'
'
surface of genus k.
Fig. 2.3.45
Suppose now that there exists at least one non-orientable strip L.
Then, as above, the part of the circle aD, singled out in Fig. 2.3.47, can
be freed from the bases of the other strips thus separating the strip L
from the other strips.
: .... :........ :.. ~.:.: .':.".

Such transformations lead to the case of several individual non-


orientable strips and several individual pairs of crossed strips. Figure ......
.-.Qn.' : .... '..
" ~.

2.3.48 shows how (in the presence of a non-orientable strip) a pair of " .': . : .'
crossed strips can be replaced by a pair of individual (non-crossed) non- " .: .....
orientable strips. Fig. 2.3.46

Thus, a manifold is obtained from a disc D by glueing several (say,


k) individual non-orientable strips with a subsequent glueing of a patch.
It is readily seen that in this case the manifold is homeomorphic to a
sphere with k Mobius strips. This completes the proof of Theorem 2.

Fig. 2.3.47

Fig. 2.3.48
128 2. Low-Dimensional Manifolds

Note, that all the 2-manifolds mentioned in Theorem 2 are distinct.


Indeed, the manifolds of the first series are orientable (i.e. contain no
Mobius strips), while those of the second series are non-orientable (i.e.
contain Mobius strips). This differs manifolds of the first series from
those of the second series since the property that "the manifold contains
a Mobius strip" is an invariant of homeomorphism. Within one series,
manifolds can be distinguished using Euler characteristics: x(8K n) =
2 - 2n and X(Mn) = 2 - n, where Kn is a solid pretzel of genus n and
Mn is a sphere with n Mobius strips.

From Theorem 2 one can also deduce the' classification theorem for
compact 2-manifolds with boundary: each such manifold M is obtained
via removal of several open discs either from the suiface of a pretzel of
some genus or from a sphere with several Mobius strips. Indeed, if we
glue components of the boundary of the manifold M with discs, we come
to a closed manifold M J , i.e. a pretzel surface or a sphere with Mobius
strips. Inversely, M is obtained from M J by removal of the glued discs.
There is another noteworthy way of specifying 2-manifolds. We take
several polygons, divide their edges into pairs and glue the edges of each
pair in one of the two possible ways. As a result, we obtain a closed 2-
manifold. Any closed 2-manifold can be obtained in this way.

2.3.6 Inversion of a Two-Dimensional Sphere

In Sect. 2.3.2 we have discussed the problem of inversion of a a circle


in a plane. This problem cannot be solved in the class of smooth im-
mersions. It turns out that the situation changes for higher dimensions:
Inverting of a a two-dimensional sphere can be inverted in R3 (in the class of smooth
2-sphere in 3-space immersions). One should not think, however, that a further increase of
dimensions retains this effect. A three-dimensional sphere cannot be
turned inside out in R4. Neither can it be done with a sphere 84 in R5
or with a sphere 85 in R6. But a sphere 86 can again be inverted in R7.
And this is the last dimension in which this operation is possible [37].
2.3 Visual Properties of Two-Dimensional Manifolds 129

All the other spheres (of higher dimensions, i.e. with n ~ 7) cannot be
inverted in ]Rn+1 (S. Smale). So, a sphere 8 n can be inverted in ]Rn+l only
for n = 2, 6. The visual inversion of a sphere 8 2 in ]R3 was constructed Visual inversion
by Arnold Shapiro [39]. See also the papers of S. Smale, N. Kuiper, B.
Morin.
Inverting 8 2 in ]R3 is rather nontrivial. We can try to draw it (see
Sect. 2.3.7). Here we shall briefly describe the main idea of this defor-
mation.
It turns out that there exists a smooth deformation of a sphere in
the immersion class which carries it to a neighbourhood of an immersed
projective plane in ]R3, namely, to a neighbourhood of the Boy surface Boy surface
constructed in Sect. 2.3.3 above. As a result, the sphere 8 2 is positioned
in ]R3 (as an immersed surface) in such a manner that it is the boundary
of the normal tubular neighbourhood of the Boy surface (i.e. immersed 2-sphere can be immersed
]R2). This means that a normal segment reconstructed at an arbitrary point in 3-space as the boundary
from JRp2 meets the immersed sphere at two close points (on both sides of tubular neighbourhood
of JRp2). We take these two points and start moving them along the of projective plane
segment towards each other. Then they meet at a point on the projective
plane (in the centre of the normal segment), then pass right through
each other and, therefore, exchange places, i.e. moving further the first
point will take the place of the second and vice versa. As a result of
this smooth deformation of the sphere its points exchange places: the
exterior point takes the place of the interior one and vice versa. Thus,
we have made the deformation of the immersed sphere (near the Boy
surface) after which the exterior and interior of the sphere exchanges
places, i.e. the sphere was turned inside out. It remains now to apply
the process, inverse to the initial one, which carried the sphere from its
standard position into the neighbourhood of the Boy surface. As a result,
we return to the standard embedding of a sphere, but now its exterior
and interior exchanged places. It the outside of the sphere was white and
the inside black, we now have a black sphere with white inside.
130 2. Low-Dimensional Manifolds

2.3.7 Visual Material

Homeomorphisms We have already discussed homeomorphisms of a sphere. Generally, to


of a sphere recognize that a given manifold is a sphere is always possible using the
algorithmical methods described above. An infinite sequence of spheres
subjected to homeomorphisms is depiced in Fig. 2.3.49. If we imagine a
sphere made of rubber, we can crumple it, expand and contract (as shown
in Fig. 2.3.49) without changing its basic topological property, the prop-
erty of being simply-connected. This means that any closed path (a loop)
2-sphere is unique shrinks along the sphere to a point. No other (closed) 2-manifold pos-
simply-connected sesses this property. Figures 2.3.50,51,52 present deformations of two-
closed 2-surface dimensional surfaces via various homeomorphisms. For convenience of
perception we use habitual images to emphasize the character (and pos-
sibilities) of such deformations. The homeomorphisms in Fig. 2.3.51 do
not change the topology of the surface and change little its area, although
they have a substantial effect upon the metric properties of the object.
"Growing" long and thin tubes-springs from the surface may, for exam-
ple, lead to the fact that the sequence of closed surfaces whose area tends
to the minimum (e.g. in a given homology class) can nonetheless sweep
Thin whiskers on a surface up the whole of the enveloping space at the expense of thin whiskers
can change its area which have no effect upon the convergence of the area to the minimum.
only a little

Page 131, Fig. 2.3.49. Group of homeomorphisms of a standard sphere.


Ancient Russian legends

Page 132, Fig.2.3.50. Two-dimensional surfaces of a high genus and


their homeomorphisms. Ancient stone temple on a sea coast

Page 133, Fig. 2.3.51. Illustration of the minimization process in the theory
of minimal surfaces. German and Scandinavian myths about the fal/ of
Valhal/a; death of the Gods

Page 134, Fig. 2.3.52. Homeomorphisms of algebraic surfaces; pOlyno-


mial of the high degree. Mystical dance of an ancient God (India)
2.4 Cohomology Groups and Differential Forms 135

2.4 What Distinguishes Two-Dimensional Manifolds? Topological invariants


Cohomology Groups and Differential Forms of 2-surfaces

As we have seen, 2-manifolds are divided into two series: spheres with Two series of 2-surfaces:
handles (they are orientable) and spheres with Mobius strips (they are orientable and non-orientable
non-orientable). Since homeomorphic manifolds should have the same
type of orientation, it follows that a sphere with handles cannot be home-
omorphic to a sphere with Mobius strips. Hence, manifolds from distinct
series differ by their orientation. We now consider manifolds of one se-
ries, for instance, spheres with handles. It turns out that spheres with
different number of handles are not homeomorphic or even homotopy
equivalent. How can this be proved? Why is, for instace, a torus not
homeomorphic and not homotopy equivalent to a sphere or, say, to a
pretzel?
The answer to this question is not an easy one. We have to find
some characteristic properties which differ a torus from a pretzel, i.e.
to discover some invariants of 2-manifolds distinct for a torus and a
pretzel. These must be invariants under homotopy equivalence. As we
already know, such invariants do exist - these are homology groups of
polyhedra. In the case of manifolds one can point out another approach to
the definition of homology groups, which we shall now briefly discuss.

2.4.1 Differentiall-Forms on a Smooth Manifold Differential 1-forms


on n-manifolds
Let M n be a smooth manifold and let Xl, . .. , xn be local regular co-
ordinates on it. We denote the differentials of these coordinates by
dxl, ... , dxn. Let the following sum, the differential expression w =
Pldx l + ... + Pndxn, where Pi = Pi(X I , ••• ,xn) are smooth functions of
the coordinates xl, ... , xn , be given in each coordinate system. We now
require that this differential expression w = E Pidxi remain invariant
under an arbitrary regular coordinate change, i.e. we require that under
transition from the coordinates xl, ... , xn to the coordinates XlI, ... ,xnl
there hold the identity E Pil dx il = E Pi dxi. Clearly, this requirement Tensor transformation law
imposes some conditions upon the transformation law of the functions for the coefficients of
Pi ---t Pil under the change of coordinates. Indeed, the rule of differen- a differential 1-form
tiation of composite functions implies
136 2. Low-Dimensional Manifolds

Since the differentials dx l ', ... ,dxn' are independent, equating the co-
efficients before them we obtain the equalities: Pi' = l:i Pi8x i/ 8xi'.
Thus, we have deduced the transformation law of the coefficients I{
in the sum w which guarantees invariance of the differential expression
w = l: Pidx i under coordinate changes.

Definition 1. The differential expression w = l: Pidx i, where the coef-


ficients PI, ... ,Pn transform under coordinate changes in the way indi-
Differential1-form cated above, is called the differentialljorm or the exterior Ijorm.
or exterior 1-form
Consider an important example of differential I-forms. Let f be a
smooth function on an n-dimensional manifold. Consider its total dif-
ferential dj. In local coordinates xl, ... , xn its explicit expression is:
df = l:(8f /8x i )dxi . Under the coordinate changes (x) -+ (x') we have
Pi' = 8f /8x i ' = l:(8f /8x i )8x i /8x i ' = l: Pi8xi /8x i ', i.e. the total dif-
Differential of a function ferential df is a differential I-form in the sense of the preceding defini-
and gradient of a function tion. It is clear that in this case the transformation law of the coefficients
Pi = 8f / 8xi guarantees the well-known invariance of first differential
under coordinate changes.

Exact differentia/Horm Definition 2. Differential I-forms w of the form dj are called exact
forms.

2.4.2 Closed and Exact Forms on a Two-Dimensional Manifold

It turns out that on a set of differential I-forms one can define a natural
Exterior differentiation operation called exterior differentiation and denoted by the same symbol
d as the ordinary differential of a smooth function. We shall not give here
the general definition of the operation d but restrict our consideration to
the concept of closed forms defined as forms whose exterior differential
is identically zero. The concrete expression of this condition is defmed
as follows.
2.4 Cohomology Groups and Differential Forms 137

Definition 3. The differentiall-fonn w = 2: Pidx i is called closed if for Closed differential forms
all i, j there hold the identities: 8Pi/8xj = 8Pj/8x i.

The reader acquainted with the general concept of exterior differen-


tiation d (e.g. from Ref. [27]) will immediately see that these conditions
are equivalent to the equality dw = O.

Proposition. Each exact form is closed. Exact forms are closed

Indeed, an exact 1-fonn has the fonn of the total differential of a


function, i.e. w = l:(8j /8x i )dx i . Then

as required. We have used the well-known fact that partial differentiations


8/8xi and 8/ 8x j always commute for smooth functions.
Consider a smooth manifold M and closed and exact 1-fonns on this
manifold. Clearly, they fonn linear spaces: linear combinations (with
constant coefficients) of exact fonns again give exact fonns. Closed 1-cocycles
fonns possess a similar property. We denote the linear space of all closed
fonns by Zl(M) and the linear space of all exact fonns by B1(M). Since
each exact fonn is closed, we obviously have an inclusion Bl(M) c
Zl(M). These spaces can be regarded as Abelian groups (by summation),
and therefore one can naturally define the quotient group Zl / Bl (or the 1-coboundaries
quotient space).

Definition 4. The quotient group Zl / Bl is denoted by Hl(M, JR) and 1-dimensional cohomology
referred to as the one-dimensional real cohomology group of the manifold group in terms of
M. differential forms

The letter JR indicates here that we are dealing with fonns with real
coefficients. Closed fonns are occasionally called cocycles and exact Cocycles
fonn coboundaries. Omitting the notation JR for the field of real numbers, Coboundaries
we shall sometimes write simply H1(M).
Using the same scheme, one can also define k-dimensional cohomo-
logy groups of a manifold, but we do not need them now and therefore
do not dwell them.
138 2. Low-Dimensional Manifolds

2.4.3 An Important Property of Cohomology Groups

So, for each manifold M we can calculate its one-dimensional coho-


mology group HI (M). Note that when defining this group we wuld
also consider manifolds with boundary but not only closed manifolds.
Consider now an arbitrary continuous deformation of a manifold. More
Homotopy equivalence precisely, consider a manifold M' homotopy equivalent to the manifold
M. Outwardly these manifolds may differ considerably. For example, a
flat ring is homotopy equivalent to a circle. A ring shrinks onto its e.g.
interior boundary which is a circle (Fig. 2.4.1). The homotopy equivalent
manifolds are here of different dimensions. Next, a Mobius strip is also
homotopy equivalent to a circle (Fig. 2.4.1). Accordingly, a flat ring is
homotopy equivalent to a Mobius strip, the ring being an orientable and
the Mobius strip a non-orientable manifold. Therefore, some essential
properties of a manifold (dimension, orientability or non-orientability)
may be destroyed under homotopy euqivalence. Homeomorphism does
not affect the dimension and orientation of manifolds. Thus, we have
Fig. 2.4.1
made sure that two manifolds can be homotopy equivalent but not home-
omorphic.
Nonetheless, the flat ring and the Mobius strip have "something in
common"; they are continuously contracted onto one and the same circle.
This "something in common" appears to consist in identical cohomology
groups.

Theorem 1. Let M and M' be homotopy equivalent manifolds (for ex-


Homotopy invariance ample, homeomorphic). Then their one-dimensional cohomology groups
of cohomology groups are isomorphic, i.e. Hl(M) = HI (M').

This property is sometimes referred to as homotopy invariance of


cohomology groups. Theorem 1 is valid for cohomology groups of any
dimension. We shall not prove here Theorem 1 since it is non-trivial
[40]. We shall only make some comments and consider corollaries. We
emphasize the importance of Theorem 1. It points to the invariants of
manifolds, i.e. objects preserved under continuous deformations - homo-
topy equivalence. What is the importance of these (and like) invariants?
They permit, for example, the proof of the fact that two manifolds are
"distinct", say, homotopically non-equivalent (and, therefore, not homeo-
2.4 Cohomology Groups and Differential Forms 139

morphic). Indeed, if the cohomology groups of two manifolds are not


isomorphic, then by virtue of Theorem 1 these manifolds are homotopi- Manifolds with different
cally non-equivalent. That this criterion is efficient we shall see below cohomology groups are
on concrete examples. homotopically non-equivalent
For manifolds there holds an important theorem: real cohomology
groups (in terms offorms) are isomorphic to real homology groups. The
latter are calculated as in the general case of polyhedra (with coeffi-
cients in the group of real numbers). Homotopy invariance of cohomol-
ogy groups can therefore be deduced from the corresponding theorem for
homologies of polyhedra. The importance of cohomology groups is ex-
plained by the fact that in many cases they are easier calculated than the
cohomology groups of a manifold (in terms of simplexes etc.). For this
reason we shall give special attention to cohomology groups. By means
of cohomology groups we shall prove, for instance, that two-dimensional
spheres with a different number of handles are not homotopy equivalent.

2.4.4 Direct Calculation of One-Dimensional Cohomology Groups How to calculate 1-dimensional


of One-Dimensional Manifolds cohomology groups?

Consider a connected I-manifold. There are only two such manifolds,


namely, the circle and the straight line.

Theorem 2. The one-dimensional cohomology group of a circle is iso-


morphic to the group of real numbers R and the cohomology group of
a straight line is zero, i.e. HI(SI) =]RI, HI(RI) = O.

Proof. We shall begin with the simpler case of a straight line. By Cohomology groups
definition, HI = Z 1/B I . We should therefore calculate the groups Z I of 1-dimensional manifolds
and BI. From the defmition of differential forms it immediately follows
that the I-form on a straight line is written as g(x)dx = w, where x is
a Cartesian coordinate on the straight line and 9 is an arbitrary smooth The case of
function. Turning to the definition of a closed I-form we see that any such a straight line
I-form on a straight line is closed because there is only one coordinate x
here, and therefore og/ox = og/ox for any function g. Thus, any Ijorm
W = g(x)dx on a straight line is closed. It turns out that any Ijorm on a
straight line is exact. Indeed, for any smooth function 9 there holds the
140 2. Low-Dimensional Manifolds

J;
identity: w = g(x)dx = d U; g(t)dt) , i.e. w = dj, where f(x) = g(t)dt.
The integral is known to depend smoothly on its upper limit, and therefore
f is always a smooth function. Thus, we have proved that on a straight
line any closed I-form is exact and, accordingly, Zl = BI. Hence the
group HI is zero. This proves the theorem for the straight line.

The case of We now proceed to the circle. On a circle we introduce the angular
a circle coordinate cp, where 0 ::;; cp ::;; 271". Then the functions 9 on the circle are
functions of the parameter cp, such that g(cp) = g(cp + 271"), Le. periodic
with respect to cp. Consequently, the general form of the I-form on the
circle is w = g(cp)dcp, where g(O) = g(271"), (Le. 9 is a periodic function in
Any 1-form on 1-manifold cp). As in the case ofa straight line, we obtain that any smooth ljorm on
is closed but not a circle is closed. It remains to calculate the group of exact forms. We
necessary exact consider the following linear map A of the space Z I of closed I-forms
into the group I. of real numbers, where A(W) = A(g(cp)dcp) = J~7r g(cp)dcp.
In other words, we associate with the form on the circle its integral along
the whole circle (Le. from 0 to 271"). Clearly, Ais linear since the integral
is additive. Hence, A is a homomorphism on the space Zl into a real
straight line 1.1.

Lemma 1. The map Ais an epimorphism, i.e. an "onto" map, its image
covers the whole straight line.

Proof. Consider the simplest I-form on a circle, namely, w = dcp.


Clearly, under the map A it will be carried into the number 27r since
A(dcp) = Jr dcp = 271". Thus, A(a dcp) = 271"a, where a = const. Accord-
ingly, the whole straight line 1.1 belongs to the image of the map since
when we change a arbitrarily we force 271"a run through all the real
numbers.

It is a well-known fact from elementary algebra that if A: B -+ A


is an epimorphism of Abelian groups (e.g. linear spaces) then we always
have an isomorphism A = B / ker A, where ker A is the kernel of the
homomorphism A, i.e. the set of all elements b E B for which A(b) = 0,
and B / ker Adenotes the quotient group (quotient space). In our case we
obtain Zl/ ker A=1.1• It remains to calculate the kernel of the map A.
2.4 Cohomology Groups and Differential Forms 141

Lemma 2. The kernel of the map A coincides with the space BI of all
exact 110rms on the circle.

Proof. We first prove that ker A:::l BI. Let w = df, where f is a smooth
function on the circle (i.e. 21l'-periodic function of 'P; f(O) = f(21l')). Then
Aw = Ii1l'df('P) = f(21l')- f(O) = 0, i.e. A(W) = 0, that is, Wbelongs to the
kernel of the map. We shall prove the inverse inclusion, namely ker AC
BI. Suppose w = g('P)d'P is a certain I-form on a circle and A(W) = 0, i.e.
Ii1l' g('P)d('P) = O. On the circle we define the function f('P) = It d(t)dt.
We shall verify whether it is a all defined function on the circle, i.e.
whether it is 21l'-periodic. We have: f(21l') = Ii1l' g(t)dt = 0 = f(O).
Here we have made use of the fact that the I-form W belongs to the
kernel of A. Consequently, we have specified on the circle a 21l'-periodic
function f('P) such that w = df since df = d ut g(t)dt) = g('P)d'P. Thus, Fig. 2.4.2
BI = ker A. Therefore, Zl / BI =]RI, i.e. HI =]RI, as required.

From Theorem 1 we iffi!I1ediately have that the circle and the straight Circle and straight line
line are not homotopy equivalent. This can however be readily seen from are not homotopy equivalent
other considerations. Note that a straight line is homotopy equivalent to a
point (Fig. 2.4.2). Hence, a circle is not homotopy equivalent to a point.

2.4.5 Direct Calculation of One-Dimensional Cohomology Groups 1-cohomology groups


of a Plane, a Two-Dimensional Sphere and a Torus of a simplest 2-surfaces

We proceed to 2-manifolds.

Theorem 3. One-dimensional cohomology groups of a plane and a


sphere are zero (trivial), and in the case of a torus T2 we have:
HI(T2) =]R2.

To begin with we calculate one-dimensional cohomology groups.


We find the general form of a closed I-form on a plane. For the general Any closed 1-form
I-form we have the expression w = P dx + Qdy, where P( x, y) and on a plane is exact
Q(x, y) are smooth functions of the Cartesian coordinates x and y. The
form is closed if and only if aP/ ay = aQ / ax (see the definition of the
closed form above).
142 2. Low-Dimensional Manifolds

Lemma 3. Any closed smooth 110rm on a plane is exact.

Different forms for the Proof. We must have a function f on the plane, such that there holds the
potential function ofa identity w = df, i.e. Pdx+Qdy = df(x,y) = (8fI8x)dx +(8fI8y) . dy.
closed 1-form on a plane This implies the following equations for the function f: 8f 18x = P,
8fl8y = Q, where 8PI8y = 8Q18x. We now integrate the first of
these equations to obtain f(x, y) = It P(x, y)dx + g(y), where g(y) is
a smooth function of one variable y. We have to find this function. To
do so, we differentiate the previous relation with respect to y and get
8fl8y = It(8PI8y)dx+8g18y. Since 8PI8y = 8QI8x and 8J18y =
0, it follows that

Q( y) = (X 8Q(x, y)d + dg(y)


x, 10· 8x x dy ,

that is,
Q(x, y) = Q(x, y) - Q(O, y) + d~~) ,

whence dg(y)ldy = Q(O,y), i.e. g(y) = I~Q(O,y)dy+C, where C is


an arbitrary constant. Thus, we obtain the explicit expression for the
unknown function f(x, y) = It P(x, y)dx + I~ Q(O, y) . dy + C. So, we
have presented a function f such that w =df, which proves Lemma 3.

By virtue of the symmetry of our arguments about the change of


x by y, we also come to the second expression for the same function
f(x, y) = I~ Q(x, y)dy + It P(x, O)dx +C'. To this end we should only
start integration with the equation 8J18y = Q. Since Zl = BI, we have
HI = 0, as required. This proves Theorem 3 for the plane.
Differential 1-form as a In so far as we are considering the plane in Cartesian coordinates,
vector field (with respect we may interpret the I-form w as a vector field v with components P
to cartesian coordinates) and Q, i.e. v = (P, Q). Then Lemma 3 will be reformulated as follows:
any closed I-form w = Pdx + Qdy on a plane determines the vector
field v = (P, Q) which is the gradient of a certain function f, i.e. v =
Potential function grad f, P = 8JI 8x, Q = 8JI 8y. Such a function is called a potential
Potential field and the field v a potential field. Therefore (in Cartesian coordinates),
closedness of the 110rm w on a plane is equivalent to potentiality of the
vector field v = (P, Q) (and to exactness of the 110rm w).
2.4 Cohomology Groups and Differential Forms 143

For what follows it would be instructive to discuss another intelJlre- 11AJit'J'(t)).!.~B


tation ofthe potential f associated with the closed I-form on a plane. To ~ '"
begin with we recall the concept of the integral of the I-form along the
curve 'Y. We set the curve 'Y as 'Y(t) = (x(t), yet)), where t is a parameter Fig. 2.4.3
(time) along the curve (Fig. 2.4.3). If on a plane the I-form Pdx + Qdy
is given, one can define its integral along the curve 'Y (e.g. from a point Integral of a 1-form
A on the curve to a point B on the curve) to be along the curve

B
r w = r P(x(t), y(t)dx(t) +Q(x(t), yet)~ dy(t) = it'r (Px +QiJ) dt ,
til til

iA it'

W
where A = 'Y(t), B = 'Y(t"), x = dx/dt, iJ = dy/dt. The same integral can
be written compactly as I"(w = I(v, "y)dt, where v = (P,Q),"y = (x,iJ)
are two vector fields along the curve (Fig. 2.4.4). Given this, the field v
depicts the form w, the field 'Y is the velocity vector field of the curve
Fig. 2.4.4
and (v, "y) denotes their Euclidean scalar product.
In these terms, the explicit formula for the potential of a closed
I-form on a plane can be rewritten in a more visual form. Consider }
on a plane a piecewise smooth path 'Y1 (Fig.2.4.5) consisting of two A Jt C
(0,1) (:IC,~)
segments - up the y-axis and then to the right, parallel to the x-axis.
j~ j"
This is the path from the point 0 (the origin) to the point (x, y). Then
the first formula for the potential f will be rewritten as f = 1"(1 w since J'2. B a::
If
f = It Q(O, y)dy + It P(x, y)dx = loA Pdx + Qdy + Pdx + Qdy =
0
Fig. 2.4.5
(:r,O)

1"(1 Pdx + Qdy. Here we assume for simplicity that f(O,O) = 0 and,
besides, that dx = 0 along the segment OA of the curve 'Y1 and dy = 0
along the segment AC of the curve 'YI.
We deduce quite similarly that there exists one more representation Visual representation
of the potential f : f = 1"(2 w, where the path 'Y2 consists of the following of a potential function
two segments: OB along the x-axis and then BC parallel to the y-axis
(Fig. 2.4.5). To integrate the values of the potential f at some point (x, y),
one should integrate the form w along a broken path starting at the point Integral of a closed form
o (say, the origin) and ending at the point (x, y). along the curves

Lemma 4. The result of integration of a closed 110rm along a path on 1-form is closed if
a plane depends only on the position of the initial and final points of corresponding vector field
the integration path and does not depend on the choice of the piecewise is potential on the plane
smooth path joining these two points. In particular, the integral of a
144 2. Low-Dimensional Manifolds

closed l{orm remains unchanged under a continuous deformation of


the path which preserves its initial and final points.

Corollary. The integral 17 w of a closed l{orm w along any closed path


(i.e. along a loop) on a plane is equal to zero.

The proof of the corollary is obvious. Inversely, if the integral of a


closed I-form along any closed path is known to be equal to zero, then
the result of integration of this form along an open path (i.e. along a path
with non-coincident initial and final points) does not depend on the path
shape but depends only on the beginning and the end of the path.
The proof of Lemma 4 follows from the fact that 171 w = 171 w for the
two paths shown in Fig. 2.4.5 as well as from the fact that any piecewise
smooth path joining the points (0, 0) and (x, y) can be approximated by
a piecewise smooth step-like path (Fig. 2.4.6) composed of the sides of
Fig. 2.4.6 rectangles as shown in Fig. 2.4.5.
Changing the initial point (the beginning of the integration path), we
change the potential f by a constant C (see above, the explicit formulas
for the potential). From Fig. 2.4.7 one can see that f = 17 W = IT+"y' W =
IT W +17' w = const+ 1', where const = IT w.
Thus, the l{orm Pdx +Qdy is closed on a plane if and only if the
Hydrodynamic interpretation corresponding vector field v = (P, Q) is potential, i.e. v = gradf. We
of a vector field shall point out a useful hydrodynamic interpretation of the condition of
closedness of the I-form won the plane.

oc:!;;>C
'Z" 0' tr
The vector field on a manifold can be interpreted as the velocity field
of a liquid flow through the manifold. Given this, the tangent vector v
fixed at a point x of the manifold should be assumed to represent the
Fig. 2.4.7 velocity vector of a particle of the liquid which, at a given moment, is
at the point x. This interpretation is useful in many respects.

Vortex-free vector field Definition 5. A liquid flow (a vector field) v = (P, Q) is called vortex-
free if oP/oy = oQ/ox.

Clearly, this condition is equivalent to closedness of the form w.


Taking together all the results obtained, we formulate the following the-
orem.
2.4 Cohomology Groups and Differential Fonns 145

Theorem 4. The following conditions are equivalent: 1) The differential Three equivalent
110rm W = Pdx+Qdy on a plane is closed. 2) The corresponding vector conditions for closedness
field v = (P, Q) is vortex-free. 3) The vector field v = (P, Q) is potential, of 1-form on plane
i. e. v = grad f. 4) The form w is exact, i. e. is the total differential of the
function f.

We now proceed to calculation of one-dimensional cohomologies of 1-cohomology groups


a sphere. This case reduces to the case of a plane. We fIx on a sphere of a sphere
the initial point 0 and suppose that we are given a closed I-form w. We
consider an arbitrary point C on the sphere and join it with the point
o by a piecewise smooth path 'Y. We state that any closed 110rm on
the sphere is exact. We present in an explicit form a function j on the
sphere, such that w = df. We assume by defInition that f"/ w = j, i.e. we
act by the scheme developed above. It is required only to prove that j
is a well defIned function, i.e. that its defInition does not depend on the
choice of the path going from the point 0 to the point C. Indeed, let 'Y'
be another such path (Fig.2.4.8). If the paths 'Y and 'Y' intersect at the
initial and fInal points only, then we may assume that between these two
paths there is a two-dimensional disc. The union of paths forms the disc
boundary. If the paths intersect (Fig. 2.4.8), then one should consider not
a single disc but rather several discs. We shall use the above arguments.
Since the paths are piecewise smooth, the disc does not cover the entire
sphere. Consequently, from the sphere one can puncture a point lying
outside the disc. As a result, the sphere will unfold to become a flat
disc, and the problem is reduced to the case of a plane. We have proved
that the integral of a closed form along any closed path on a sphere is
equal to zero. Thus, the function (potential) j is well defIned on the
sphere. We have made use of the fact that on a sphere any two paths
with coincident fInal and initial points can be continuously deformed into
each other (without displacing their end-points). On the other hand, it is
clear that w = dj. We have presented a smooth single-valued (on the Fig. 2.4.8
sphere) potential and proved exactness of the form. Hence, HI(SI) = O.
Let us consider a torus. It is again convenient to represent it as a
square with identifIed sides (Fig.2.4.9). On this torus (square) we in-
troduce Cartesian coordinates x, y and set two I-forms WI = dx and
W2 = dy (with constant coefficients). They can be represented (following
the above principle of association of the I-form with the vector fIeld)
146 2. Low-Dimensional Manifolds

as fields VI = (1,0), V2 = (0, 1). Both the fields are shown in Fig. 2.4.9.
Calculation They represent liquid flows along the x- and y-axes respectively, having
of 1-cohomology constant velocities equal to 1.

:··,s:···1
group of a torus Let Zl be the linear space of closed I-forms on a torus. Consider

E:j
a linear map (homomorphism) ,\ of the group Zl into an Abelian group
]R2, i.e. into a Euclidean plane: '\(w) = (a,{3), where a = Ia w,{3 = Ibw
.' .
. .. ',' are two numbers which are respectively the integrals of the I-forms w
a'.:'. a along two circles a and b on the torus, namely, along the meridian a
" ..... ~: ::- and the parallel b. When a torus is made of a square, the side a of the
11',f, square becomes the circle-meridian and the side b the circle-parallel. In
the explicit form, ,\ will be written as follows. The general form of the


I-form on the torus is Pdx +Qdy, where P and Q are doubly periodic
functions on the plane (i.e. periodic in x and y). Then

,\(Pdx + Qdy) = (l P(x, O)dx, l Q(O, Y)dY)

We state that ,\ is the map onto the entire plane, i.e. ,\ is an epi-
morphism. Indeed, if as the I-form w we take WI = dx, we obtain
'\(dx) = (1,0). Taking as w the form W2 = dy, we have '\(dy) = (0, 1).
Since the vectors (1,0) and (0, 1) form the basis on the plane, ,\ is an
epimorphism, as required. Both the forms dx and dy are closed on the
torus, i.e. belong to the space Zl. Consequently, from the general al-
gebraic considerations it follows (as in the case of a plane or a sphere)
that Z 1/ ker ,\ = ]R2. It remains to prove that ker ,\ = B I, i.e. that the
kernel of >. coincides with the space of exact I-forms. If the form is

° °
exact, i.e. w = df, where f is the function defined on the entire torus,
Fig. 2.4.9 then lad! = and Ibdf = since a and b are closed paths (cycles).
Therefore, BI C ker >.. We shall prove inversely that BI :J ker >.. Let
Parallels and meridians the closed I-form w be such that its integrals along the meridian and
of a torus as the 1-cycles parallel on the torus are equal to zero. It should be proved that the form
Parallel and meridian is exact, i.e. that there exists a potential f on the torus, such that w = df.
form the basis on a torus We shall present it in the explicit form. We set by defmition f = I"(w,
Closed 1-forms where 'Y is any piecewise smooth path going from a fixed point 0 on
on the torus the torus into a variable point (x, y) on the torus. It should again be
proved that this defmition is independent of the choice of the path going
from the point (0,0) to the point (x, y). As we know, this will follow
from the fact that the integral of a given closed form along any closed
2.4 Cohomology Groups and Differential Forms 147

path (loop) is equal to zero. Let 'Y be any such path on the toms. As is
already known, a continuous deformation of a path leaves the integral
of a closed I-form along this path unchanged. Next, it is geometrically
obvious that a path can always be deformed into a path moving along a
certain parallel and a certain meridian of the toms (in an arbitrary order).
For example, the path will go alon~ the parallel, then several times along
the meridian, then again along the parallel, etc. Consequently, the initial
integral of the I-form along the path 'Y is equal to the integral of the
same form but along a new path (which is described above) composed
of several copies of the parallel and several copies of the meridian (per-
haps with reverse orientations). By virtue of additivity of the integral,
the initial integral II w is represented as the sum of the integrals of the
form w along the parallel and meridian, and each of these integrals is,
generally, taken several times according to how many times the path 'Y
revolved along the parallel and how many times along the meridian (with
allowance made for the direction of revolution). But since each of these
integrals is equal to zero, so is the initial integral. We have proved that
the function f = II W is well defined on the entire toms. Next, as in the
planar case we obtain that w = df. Consequently, a closed 110nn with
zero integrals along the parallel and meridian of the toms is an exact 1-cohomology group of
form. Thus, BI = ker >., whence HI (T2) =]R2. a torus is isomorphic to 1.2
So, we begin revealing the following regularity. A sphere is a closed
manifold without handles and has a zero group HI. A toms is a sphere
with one handle and has a nonzero group HI isomorphic to ]R2. Thus,
glueing one handle to a sphere we increase the one-dimensional coho-
mology group by the group ]R2. It is natural to assume that if we glue
one more handle to the sphere (i.e. obtain a pretzel, a sphere with two
handles), the cohomology group will again increase by]R2, and as a result
the one-dimensional cohomology group of the pretzel will be isomorphic
to ]R4.
In particular, we have in fact proved the theorem: a closed 110rm Exact 1-forms on a torus
on a torus is exact if and only if its integrals along two closed circles and zero periods of the
- the parallel and meridian - are equal to zero. The extension of this form along parallel
criterion to the case of a sphere with an arbitrary number of handles is and meridian
discussed in the subsection to follow.
148 2. Low-Dimensional Manifolds

1-cohomology groups 2.4.6 Direct Calculation of One-Dimensional Cohomology Groups


of oriented 2-surfaces of Oriented Surfaces, i.e. Spheres with Handles

Theorem S. Let M2 be a closed compact connected orientable 2-


manifold, i.e. a sphere with 9 handles. Then its one-dimensional co-
homology group HI is isomorphic to ]R2g.

The proof will be carried out by the scheme already elaborated for
the plane, sphere and torus. The crucial point of this construction is to
present in an explicit form a sufficiently large set of independent closed
I-forms.
We consider a surface of genus 9 and represent it as a convex flat
Coding of a 2-surface 4g-gon with sides aj and bj, where 1 ::; i ::; g, the code W being
of the form alblallbll ... agbgaglbgl (Fig. 2.4.10). On the surface we
construct closed I-forms WI, ... ,Wg and 71, .•• , 7g • If a closed path (such
paths are sometimes called cycles) is given on the surface, the integral of
Period of a form the form along the path is called its period along the cycle. We single out
along the cycles the cycles al, ... ,ag and bl, ... ,bg on the surface and call them basis
cycles (from the considerations described below). Each of the cycles is
represented as a circle on the surface. To each handle there correspond
exactly two such cycles: aj and bj, where i is the number of the handle
(Fig. 2.4.10). For each of the forms Wj and 7j (to be constructed just now)
its period by the basis cycles is defined. Calculating the periods, we are
Matrix of periods led to a square matrix called the period matrix. We wish to construct
I-forms Wj and 7j such that the period matrix be a unit matrix of the
form

c. :)
Let us construct the form WI. We single out four edges of the poly-
gon W corresponding to the first letters of the code W, namely, to the
commutator alblallbll. On the edge al we choose three successive
points G, B, A and then consider the corresponding points G', B', A' on
the edge all (Fig. 2.4.10). We join corresponding points with segments,
i.e. draw segments AA', BB', GG'. When the surface M is glued of the
2.4 Cohomology Groups and Differential Forms 149

fundamental polygon W, the segment eA is identified with the segment Fundamental polygon
e' A'. We now construct on the polygon the function fas follows:

1) Set f == 0 outside the rectangle eAA'e' (the area covered with


points in Fig. 2.4.10).
2) Set f == 1 on the rectangle BAA' B' (the dashed area in
Fig. 2.4.10)
3) In the rectangle eBB'e', take for f a smooth function equal to
zero on the side ee', equal to unity on the side BB' and smooth inside ' ..

e e'
the rectangle, including the sides Band B'. The restriction of the
----_ ..
function on the side eB coincides with that on e' B', and after glueing
the sides eB and e' B', we obtain a smooth function.

It should be emphasized that the function that we have constructed


on the polygon W is discontinuous! The line of its discontinuity is the
segment AA'. On one side of this segment the function is equal to unity
and on the other side to zero. As the I-form WI we take the form df.
With this formula the I-form is well defined everywhere on the polygon
except on the segment AA'. On the segment AA' we set by definition
WI = O. This yields a smooth form on the entire polygon and on the entire Fig. 2.4.10
surface obtained via glueing the sides of the polygon. Smoothness of the
form follows from the construction and from the fact that the function Construction of the basis
f is constant in a certain neighbourhood on the right and on the left of 1-forms on 2-surface
the segment AA', i.e. in this neighbourhood df = 0. Defining the form
WI, in addition, by zero on the segment AA', we just obtain the form
smooth already at all points of the surface. The form is closed. Indeed,
the condition of closedness, i.e. dwl = 0, outside the rectangle eBB'e'
is obvious, as the form is here equal to zero. In the rectangle eBB'e'
the form is written as df and, therefore, dw = d2 f = 0, as required.
Quite similarly, but now taking the other two edges, bl and bl l ,
we construct the second I-form TI on the surface. Considering the other Commutators in
commutators aibia;\-I, we construct the other forms Wi and Ti. As a fundamental polygon
result, on the surface of genus 9 we present a set of 29 closed I-forms: and geometrical 1-forms
WI,TI;W2,T2; ... ;Wg ,Tg• Soon we shall see that none of them is exact.
Although the function fi participated in their construction, it cannot be
continued to become a smooth function on the entire surface. In our
construction this function was discontinuous.
150 2. Low-Dimensional Manifolds

Space of closed 1-forms We denote the linear space of all closed I-forms on the surface by
on a 2-surface Z I and construct the map .A of this space into the space (group) 1.2g by
the formula: .A(w) = (0'.1, (31, ... , O'.g, (3g), where O'.j = fai Wi, (3i = fbi 7'i.
Here Wi and 7'i are the forms constructed above. We state that the map
.A is an epimorphism. Indeed, calculate all the periods of the forms Wi
and 7'i. Take, for instance, the form WI. Its period along the circle al is
equal to fal WI = fg WI = I(B) - I(C) = 1 - 0 = 1. At the same time,
the period of the same form along any of the basis cycles other than al
is obviously equal to zero. The point is that this form is equal to zero on
all the other basis cycles. Hence, .A(WI) = (1,0,0, ... ,0).
It can be verified quite similarly that .A(7'I) = (0,1,0, ... ,0). Finally
Period matrix we obtain that the period matrix of the forms WI, 7'1, ... , wg, 7'g along
of a basis 1-forms the cycles ai, bl, ... , ag , bg is a unit matrix, as required. Considering
on 2-surface in the space 1.2g the standard basis consisting of vectors all of whose
coordinates are zero except one (which is equal to 1), we obviously just
obtain that these basis vectors get into the image of the map .A. Therefore,
the image of the map .A covers the whole space 1.29 • As before, we shall
prove that ker .A = B I, i.e. that a closed l{orm on the surface is exact if
and only if all its periods along basis cycles are equal to zero. We shall
prove the lemma valid for any smooth manifold.

All periods of an exact Lemma 5. If the l{orm W has the form dl, where I is a smooth function
1-form are equal to zero defined on the entire manifold, the integral ~f the form along cycle 'Y is
equal to zero.

Indeed, f"{w = f"{dl = 1('Y(I» - 1("f(0» since the path 'Y is closed,
and therefore 'Y(O) = 'Y(1).
This implies that BI C ker .A. We shall prove the inverse inclusion.

If all periods of a closed Lemma 6. Let W be a certain closed I-form on M, all of whose periods
1-form are equal to zero, along the basis cycles on the surface are equal to zero. Then this form
then the form is exact is exact.
Explicit form We shall construct the potential I in a explicit form, i.e. such a
of a potential for function that W = df. We fix on the surface an arbitrary point 0 (''the ori-
the exact 1-form gin"), take an arbitrary point P and join it with a piecewise smooth path
'Y on the surface. By definition I(P) = f,,{ w. It is required to prove that
2.4 Cohomology Groups and Differential Forms 151

this function is well defined, i.e. does not depend on the choice of the
path. This will obvioulsy follow from the fact that the period of the closed
I-form w along any closed path (cycle) "( is equal to zero (in our assump-
tions). We take an arbitrary piecewise smooth closed path and subject
it to a continuous deformation. Then the integral i'"l w does not change.
Indeed, any continuous deformation of a path can be represented as a
sequence of the following simplest (elementary) deformations "( -+ "(' I
(Fig. 2.4.11). On the other hand, the path difference "( -+ "(' (Fig. 2.4.11)
may be thought of as the boundary of a two-dimensional disc. Therefore,
we may apply Lemma 4 to obtain that under continuous deformation the
/~
r-
unknown integral remains unchanged. Next, continuously deforming an Fig. 2.4.11
arbitrary closed path on a closed 2-surface (i.e. on a sphere with handles)
we can make it become a path composed of several copies of elementary
paths-cycles a" b" ... ,ag , bg (Fig. 2.4.10) taken with certain multiplic-
ities (one and the same loop can be passed several times and, perhaps,
in different directions and in an arbitrary order). This transforms the un-
known integral i'"l w into the sum of integrals of the I-form along the
loops ai and bi passed, may be, several times. But these individual in-
tegrals are equal to zero' by the assumption that all the periods of the
form along the basis cycles are zero. Consequently, the integral of the
form along any closed cycle is also zero. This completes the proof of
Lemma 6.
Since w = df, where f is the function constructed above, we have
proved exactness of any closed I-form with zero periods along basis Exactness of any closed
cycles. Hence, ker A = B' and H'(Mi) = R2g , which implies the 1-form with zero periods
theorem.

Corollary. Two-dimensional closed manifolds of different genus are not 2-surfaces of different
homeomorphic and not homotopy equivalent. genus are not homotopy
equivalent
Indeed, we have HI = R2g, where 9 is genus. If p =f q, the Abelian
groups R2p and R2q are not isomorphic. Therefore, a sphere with p han-
dles and a sphere with q handles have different cohomology groups. By
virtue of Theorem 1, these manifolds are not homotopy equivalent.
152 2. Low-Dimensional Manifolds

Algorithmical recognition 2.4.7 An Algorithm for Recognition of Two-Dimensional Manifolds.


of 2-surfaces Elements of Two-Dimensional Computer Geometry

Consider a set of closed connected compact 2-manifolds. Let us formulate


Elements of the question which is at first glance elementary and the answer to which
computer geometry seems to be obvious: What does it mean when we say "consider the set of
all two-dimensional manifolds"? What do we mean by saying "to specify
a manifold"? We are already acquainted with one of the versions of the
Codes of a manifold answer. Specifying the code word Wand the corresponding fundamental
polygon, we determine a 2-manifold by glueing the sides of the polygon.
Here we have specified the manifold by its code W. If, however, a
manifold is given in some other manner, say, by algebraic equations then
it is very difficult to establish what fundamental polygon corresponds to
this manifold. The question arises of how a two-dimensional manifold
(say, a sphere) can be identified among the variety of the manifold coded
in a definite way (generallly, in a way other than that descrived above).
As we know from Chap. 1, a convenient way of specifying (coding) 2-
Triangulation manifolds is presentation of their triangulation. Recall that any smooth
manifold can be triangulated. Let us number all the triangles, edges and
vertices of a triangulation and indicate the edges or vertices in which the
triangles are glued. Let us write, for example, that triangles 5 and 15 are
glued along edge 9 and edges 8 and 15 are g~ued at vertex 578. This will
give us a table. Chapter 1 acquainted us with one of the versions of such
Incidence matrix a table, with incidence matrices. We shall denote this resultant table by
T(M). Obviously, it allows us to restore a given 2-manifold uniquely.
So, the table T(M) can play the role of the code of a manifold. We can
specify manifolds by such tables. This way is convenient for it rests on
the simple principle: we indicate ''which is glued to which". This way
is applicable in a much more general situation than coding 2-manifolds
by means of their fundamental polygons. The point is that if a surface
is given by equations (or by some geometrical rule) then the surface can
Compatible triangulations practically always be effectively subdivided into small enough triangles,
in different portions the process beginning with a point and gradually extending to the entire
of the surface surface. Sometimes we have to go back to subdivide some of the triangles
already constructed into smaller one i.e. reduce the division in order to
make the triangulation compatible in different portions of the surface. If
2.4 Cohomology Groups and Differential Forms 153

the surface is compact, the process will stop sooner or later. It is essential
that we should work in each step only on a small piece of the surface.
Let a 2-maifold be realized in space as a smooth surface and let
a small geodesist be "landed" onto this surface with a task to draw an
accurate map of the surface and to establish its topological type. He Map of the surface
cannot see the whole surface at a time and cannot answer the question
on what precise manifold he is (Fig.2.4.12), or whether there exists an
algorithm for surface recognition. That is, he cannot. say whether it is Surface recognition
orientable or not, how many handles or Mobius strips it has. If the reader
ponders over the problem seriously, he is sure to find the answer to be
not an easy one. A simple detour round the surface will give nothing:
one does not know how to discover handles and how to count them. It is
intuitively clear that to find out what fundamental polygon this surface
has is practically impossible by way of simple observation. The formal
knowledge (which is geodesist has) of the fact that the surface has some Fig. 2.4.12
fundamental polygon is of no use here.
However, we already know that the geodesist can solve the problem ~ _ o OllC>OClQ{je •••• •
without leaving the surface. We shall describe this method once again II)fl-- ooQ'i?0'ii'oQQ .... •
giving attention to the subject matter of the recognition process. Within a ~ ¥ __ 'OQ~I>~09~· .. ..
fmite period of time, the geodesist can draw on the surface a set of trian- .!!~ ___ O'OP DDo9Q .... •
.~" ___ t3~O(l:>e.&t'c:O ... ..
gles to obtain triangulation. The work may appear to be cumbersome, but ~; codes of manifolds
moving continuously along the surface and sometimes returning he can
make all the triangles compatible. The geodesist numbers successively Fig. 2.4.13
(without turns) all the triangles, edges and vertices. The result of this
cartography will be the table T(M). The process of partition of a mani- Cartography process
fold into triangles is of course ambiguous. For example, if the geodesist
erased this triangulation and repeated the whole process agian, he would
almost sure to get the result different from the initial one. So, one and
the same manifold can be associated with an infinite number of tables
T(M) that code a given surface (Fig. 2.4.13). Thus, we can construct the
algorithm for enumerating all 2-manifolds and obtain the list of them. Algorithm for
The list will be excessive in the sense that one and the same manifold enumerating all 2-surfaces
is represented there an infinite number of times, i.e. acquires an infi-
nite number of codes. The problem arises of how we can know from
two codes in the list whether they specify one and the same or different
manifolds.
154 2. Low-Dimensional Manifolds

Algorithm for recognition Such an algorithm for recognition does exist, although in practical
and incidence matrices work it may require much time to give an answer. This necessary time can
however be efficiently estimated from above using a calculable constant
which depends only on the size of the initial table T(M). This procedure
is described below.

1-dimensional cohomology 2.4.8 Calculation of One-Dimensional Cohomologies of a Surface


groups and triangulation Using Triangulation

We should now return to the material of Chap. 1. Recall that the real
cohomology groups of smooth manifolds coincide with real simplicial
homology groups of these manifolds calculated using incidence matrices.
The problem therefore reduces to the analysis of the incidence matrices
which are in fact already known to the geodesist since he has already
compiled the table T(M) corresponding to a given triangulation. The
geodesist should calculate all the boundary operators (Fig.2.4.14) and
Fig. 2.4.14 find all the one-dimensional cycles and boundaries (Fig.2.4.15). Then,
using personal computer, he can quickly reduce the incidence matrices to

~
1. the normal form and calculate the real homology groups. Moreover, he
•. :::.L13 can determine whether the surface is orientable or not. Using the familiar
"fJ......
.. ' to him theoretical information on the structure of homology groups of
/1;" 2-manifolds, the geodesist answers the question.
Fig. 2.4.15
2.4.9 Visual Material

In Sect. 2.4 we have described only the simplest methods of calculation of


Spectral sequence method homologies and cohomologies. In addition, the spectral sequence method
is frequently employed. See e.g. Refs. [27], [24]. Figure 2.4.16 represents
the initial stage of triangulation of a two-dimensional surface, necessary
for algorithmic clarification of the topological structure of the object on
which the geodesist works (see Sect. 2.4.7). Using a "gas burner" one
can cut the surface into triangles and then glue of them a flat funda-

Page 155, Fig. 2.4.16. Triangulation and coding of 2-surfaces. Medieval


German and Scandinavian myths about Sigfried and Brunhilda (death of
Sigfriedj
156 2. Low-Dimensional Manifolds

mental polygon - the code of the surface. Figure 2.4.17 is an infmite


sequence of spheres and in the background there is a sphere subjected
Triangulation of to homeomorphism. Triangulation of a polyhedron and the process of
a polyhedron its cutting into elementary pieces, blocks are depited in Fig. 2.4.18. The
polyhedron is transformed here into a chaotic cluster of simplexes (or
parallelepipeds) each of which remembers, however, its neighbours since
the incidence matrices restore completely the polyhedron. Therefore, all
the blocks may, if necessary, again unite to form a polyhedron. The nec-
Memory of an essary memory of an individual block (simplex) is very small - it must
individual block remember only its immediate neighbours to which it is incident. Thus,
Complex system a complex system can be effectively constructed of a large number of
individual elements with a comparatively small memory.

Page 157, Fig. 2.4.17. Differential geometry and topological properties of


a smooth 2-surfaces; the main step in the proof of Gauss-Bonnet formula.
Expulsion of ancient Greek gods from Olymp

Page 158, Fig. 2.4.18. Incidence matrices determine the topological struc-
ture of a complex objects; memory af an individual block reconstructs its
neighbourhood
2.5 Visual Properties of Three-Dimensional Manifolds 159

2.5 Visual Properties of Three-Dimensional Manifolds

2.5.1 Heegaard Splittings (or Diagrams)

The problem of classification of 3-manifolds is much more complicated Classification of 3-manifolds


than that of2-manifolds and it has not yet been completely solved even in
the case of closed 3-manifolds. The solution of the classification problem
consists in the construction of two algorithms: the algorithm for enumer- Algorithm for enumeration
ation of manifolds (repetitions are admitted) and the algorithm for iden- Algorithm for identification
tification of manifolds. The latter establishes in each pair of manifolds
(codes) whether they are identical or not. There exist several good algo-
rithms for enumerating closed 3-manifolds. Each of them i~ based on a
special way of specifying 3-manifolds. A complete algorithm for iden-
tifying (recognizing) 3-manifolds has not been constructed up to now.
There does not known even an effective simple algorithm for recognizing
a standard sphere 83, alth9ugh partial algorithms, for example, in the
class of so-called 3-manifolds of genus 2 do exist. Birman and Hilden
[41] have proved algorithmic recognizability of Heegaard diagrams of the Heegaard diagrams
sphere 83 in a certain class of Heegaard diagrams of 3-manifolds, this Partial algorithms
class containing, in particular, all Heegaard diagrams of genus 2. But they
have reduced the problem to the famous Haken algorithm (recognition of Haken algorithm
a trivial knot) which cannot be practically realized using computer. So,
the problem of finding an effective algorithm admitting computer real-
ization remained urgent. This algorithm (for genus 2) was proposed and
experimentally worked out by I.A. Volodin and A.T. Fomenko in 1974
and realized using computer together with V.E. Kuznetsov (see Ref. [42])
(so-called cut-vertex algorithm). The efficiency of this algorithm was fi- Cut-vertex algorithm
nally grounded by Homma, Ochiai and Takahashi [43] in 1980. A more for recognition of
detailed and demonstrative presentation of the algorithm is given in the a standard sphere
books [28] and [29]. Here this material is omitted.
One of the most frequently used classical ways of specifying 3-
manifolds is Heegaard splittings and their modifications (Heegaard di-
agrams, nets, special splittings). For brevity, we shall henceforth call a
sphere with 9 handles a pretzel (of genus g). Embedding such a pretzel Solid pretzel of genus g
into R3, we may consider a three-dimensional manifold (body) which is
restricted by this pretzel and is called afull (or solid) pretzel of genus g.
160 2. Low-Dimensional Manifolds

Suppose there exist two copies Kg, K~ of a solid pretzel of genus


9 and a certain homeomorphism h: 8Kg --t 8K~ of their boundaries.
Glueing the pretzels Glueing the pretzels via this homeomorphism we obtain a closed 3-
manifold M. Note that the manifold M is represented as a union of the
two solid pretzels (the images of the pretzels K~ and Kg under their
glueing) lying in it.

Heegaard splitting Definition 1. The Heegaard splitting of a closed orientable 3-manifold


M is the representation of this manifold as a union of two pretzels with
common boundary, lying it it. Since the pretzel boundaries coincide, the
Genus of the pretzels are of the same genus. It is called the genus of the Heegaard
Heegaard splitting splitting.

Theorem 1. Any closed orientable 3-manifold admits Heegaard splitting


of some genus.

The proof is carried out using the same technique as in the proof
of the two-dimensional surface classification theorem. We preliminarily
introduce a very important concept of a handle.
Let M be an n-manifold and let h : 8 A- 1 X Dn-A --t 8M be a certain
embedding. Then we say that the manifold MI = M Uh(D A X Dn-A)
n-dimensional handle is obtained from the manifold M by glueing an n-dimensional handle
of index'\ of index A. The image of the manifold DA X Dn-A in the manifold MI
Base of the handle under glueing is called a handle and the manifold h(8 A- 1 X Dn - A) is
called the base of it. The disc, the strip and the patch from the proof
of Theorem 2, Sect. 2.4.3 are two-dimensional handles of indices 0, 1
and 2 respectively. The glueing of a handle of index 0 to an n-manifold
consists in the addition of an n-ball taken separately; the glueing of a
handle of index n consists in glueing up the spherical component of the
boundary by an n-ball. Examples of three-dimensional handles of indices
Fig. 2.5.1 1 and 2 are given in Fig. 2.5.1.
Splitting into An n-manifold is said to be splitted into handles if it is obtained
union of handles from a ball by a successive glueing of handles of different indices. There
exists the following duality: any handle of index A glued to the union of
Duality between the preceding handles can be regarded as a handle of index n - Aglued
the handles to the union of the remaining handles.
2.5 Visual Properties of Three-Dimensional Manifolds 161

The proof ofTheorem 1 is carried out by the following scheme. First


we prove that an arbitrary 3-manifolds can be trianglulated, i.e. parti- Triangulation
tioned into curvilinear tetrahedra in such a way that any two tetrahedra of 3-manifolds
either do not intersect at all or intersect at a vertex, along an edge or a
face. To db so, one should be able to make triangulations of Euclidean
neighbourhoods of points compatible. As was proved by Moise, this can
always be done in three dimensions. It should be noted that one cannot,
generally, achieve such a compatibility in higher dimensions.
Then the vertices, edges and triangles are replaced by handles of in-
dices 0, 1 and 2 respectively. The remaining "middle" parts of tetrahedra
are handles of index 3 (Fig. 2.5.2).
If two handles of index 0 are joined by a handle of index 1, their
union is assumed to be a new handle of index 0, which will decrease
the total number of handles. In a dual manner, two handles of index 3,
which a handle of index 2 joins with its free sides, can be united with
this latter one to form a new handle of index 3. Repeating this argument, Fig. 2.5.2
we arrive at the case of one handle of index 0 and one handle of index
3. Then the union of a handle of index 0 and handles of index 1 is a ball Ball with handles
with handles, i.e. a solid pretzel. Its complement is a union of handles
of index 2 and a handle of index 3, i.e. from the dual point of view the
union of handles of index I with a handle of index 0, which is a solid
pretzel too. As a result we obtain a Heegaard splitting.
Obviously, the only 3-manifold of genus 0 (i.e. the one obtained by
glueing two balls) is the sphere. The sphere also posseses the Heegaard
splitting of genus 1 (and the Heegaard splitting of any higher order).
This can be most easily seen from the "differentiation" formula: 8 3 =
a(D2 x D2) = aD 2 x D2 U D2 x aD 2, where aD 2 x D2 and D2 x aD2
are pretzels of genus 1 (solid tori). Identification of points of the sphere
a(D2 x D2) symmetric with respect to zero yields a three-dimensional Fig. 2.5.3
projective space lRp3. Identification of points of each of the solid tori also
yields solid tori. Thus we have constructed an example of the Heegaard Heegaard splittings
splitting of genus 1 of a manifold ]Rp3. The Heegaard splittings of a of a sphere and a
sphere and a projective space can be seen in Fig. 2.5.3 and 4. The sphere projective space
8 3 is represented in Fig. 2.5.3 as a space lR3 compactified by an "infinitely 3-sphere
remote" point. One of the tori of this splitting is obtained by rotation of
the ring A around the axis l which, together with the infinitely remote
point, composes a circle. The complement of this torus is splitted into
162 2. Low-Dimensional Manifolds

discs each of which intersects the i-axis at exactly one point and is
obtained by rotation of a corresponding "solenoidal" arc. Therefore, this
complement is also a solid torus. Note that meridians and parallels of
solid tori exchange places under glueing: a meridian becomes a parallel
and vice versa.
In Fig. 2.5.4 the space ID'3 is represented as a space JR.3 with im-
proper points added by one for each direction (the class of parallel straight
lines). The one-sheeted hyperboloid splits JR.p3 into two solid tori. The
meridional disc of one torus coincides with the disc spanned by the neck
cross-section of the hyperboloid, the meridional disc of the other one is
composed of parts (31, f32 of the plane passing through the hyperboloid
axis. As the parallel of both the tori one can take the rectilinear generator
of the hyperboloid.

2.5.2 Examples of Three-Dimensional Manifolds

Fig. 2.5.4 We are already acquainted with the simplest examples of three-dimen-
sional manifolds. This is the standard sphere 8 3 given by the equation
(xli + (x 2)2 + (x 3i + (x4)2 = I in a Euclidean space JR.4; the three-
Projective space lRp3 dimensional projective space JR.p3 whose points are, for example, straight
Direct product 8 1 x 82 lines in JR.4 passing through the origin; the direct product of a circle by a
two-dimensional sphere, i.e. 8' x 8 2 and, finany, the three-dimensional
3-torus torus T3 which is a product of three circles, i.e. 8 1 x 8 1 X 8 1. Besides, in
Lens spaces concrete applications we often encounter so-called lens spaces (or simply
"lenses") Lp,q defmed like this. We consider a sphere 8 3 embedded
standardly in JR.4 which is identified with the complex space C2(z, w).
The sphere is given here by the equation Izl2 + Iwl 2= 1. Let us consider
the action of the Abelian group Zc on the sphere 8 3, which is given by
the formula (z, w) - t (e27riP/C. z, e27riq/c • w). The linear transformation
defined by this formula obviously maps the sphere into itself Considering
the iterations of this transformation, we determine the action of the other
elements of the group Zc. Factorizing the sphere by the action of this
group (Le. identifying the points obtained from one another via group
transformations) we obtain a new space which is just called the lens
space. In a particular case, considering a lens Ll,1 for c = 2, we obtain
a projective space. In this case the action of the group Z2 is given by
0;
2.5 Visual Properties of Three-Dimensional Manifolds 163

the formula (z, w) - (-z, -w) which describes an involutive self-map


of the sphere. Each point of the sphere is mapped into a diametrically ~

,gOl
opposite one. 1 ._~
.., ......

New three-dimensional manifolds can be alternatively obtained like


this. Suppose we are given two n-manifolds Ml and M2. In each of
them we single out an n-dimensional ball D and discard it. We obtain G
~I I.....
I,
1 .....-
_
two manifolds with boundary homeomorphic to a sphere sn-l . Consider
a cylinder (a tube) Dl X sn-l, where Dl is a segment. The cylinder
boundary consists of two spheres. We glue these spheres to the boundary
spheres of the manifolds Ml \D and M2 \D (Fig. 2.5.5). We obtain a new
manifold which we denote by Ml UM2 and call it the connected sum of
the manifolds Ml and M2. Note that the angles arising in the course of
glueing the cylinder (the tube) to the holes can be smoothed, as shown
in Fig. 2.5.5, so that we again have a smooth manifold.
The operation of taking a connected sum is especially obvious on
an example of2-manifolds (Fig. 2.5.6). If we take a connected sum ofa
sphere with any manifold, we shall have nothing new: the result will be Fig. 2.5.5

0"'"
homeomorphic to the initial manifold (Fig.2.5.6), i.e. Mn Usn ~ Mn.
In particular, sn U..• usn = sn. A connected sum of two tori gives .~ '\
a pretzel. Generally, a connected sum of a 2-manifold with a torus is ,
equivalent to glueing one handle. Thus, any closed orientable connected
2-manifold, i.e. a sphere with handles results from taking a connected
sum of a corresponding number of tori. In this sense S2 and the torus
T2 are "independent" (in the class of orientable manifolds). The other
orientable manifolds are decomposed into connected sums o~ these man-
ifolds. On this basis we give the natural defmition. A manifold M n is
called irreducible (in the topological sense) if it cannot be represented
as a connected sum of two manifolds Ml and M2 distinct from the stan-
dard sphere, i.e. M f Ml UM2. The other manifolds are called reducible.
Therefore, the manifold M is reducible provided that M = Ml ~ M2,
where Mi f sn, i = 1,2. From this point of view, the standard sphere is
assumed to be an irreducible manifold (although it falls into a connected Fig. 2.5.6
sum of an arbitrary number of spheres).
164 2. Low-Dimensional Manifolds

2.5.3 Equivalence of Heegaard Splittings

Equivalent Two Heegaard splittings of a 3-manifold M are called equivalent if there


Heegaard splittings exists a homeomorphism of the manifold M onto itself which maps the
Heegaard surface (i.e. the common boundary of cracknels) of one split-
Heegaard 2-surface ting into the Heegaard surface of the other splitting. In this connection
we may present the following information.
1) Any two Heegaard splittings of a sphere, of equal genus, are
equivalent (the result of Waldhausen).
2) Any two splittings of genus 1 of a manifold of genus 1 are
equivalent.
3) There exist non-equivalent splittings of genus 2 of a connected
sum of two lens spaces.
4) Any two Heegaard splittings of one and the same manifold are
stable-equivalent, i.e. become equivalent after each of them undergoes
several stabilization operations, namely, local addition ofa trivial-handle
splitting to a pretzel (Fig. 2.5.7).

Fig. 2.5.7 Note that none of these facts is trivial. The first of them admits a
beautiful reformulation in the language of homeomorphisms of a pret-
Homeomorphism group zel surface. We denote a homeomorphism of the surface of a standard
of the pretzel pretzel, which maps its meridians into parallel~, by cpo Then the homeo-
morphism h determines a sphere if and only if h has the form acpb,
where the homeomorphisms a and b are continued onto the interior of a
solid pretzel. The structure of the homeomorphism group of the pretzel
surface is well-known. Lickorish has found the generators and Wein-
reeb the relations (see e.g. Referativnyi Zhumal (Review Journal) 1984,
Homeomorphism group 4A597). The structure of a homeomorphism group of a solid pretzel is
of a solid pretzel also well known. For the review of these results see Ref. [29]. The prob-
lem of algorithmic recognition of a sphere is reduced to the problem of
recognition of bilateral cosets of a known group by a known subgroup.
From the second class one can readily obtain the topological classi-
fication of lens spaces (without the use of Rademeister torsion).
Using the fourth fact one can reuce the problem of classification
of all 3-manifolds to the problem of classification of bilateral cosets of
a stabilized homeomorphism group of a pretzel surface by a stabilized
homeomorphism group of a solid pretzel.
2.5 Visual Properties of Three-Dimensional Manifolds 165

Let M3 be represented as a union of two solid pretzels Kg and K~


with a common boundary Hg = 8Kg = 8K;. The meridians of the pret-
zels Kg and K; form a set of curves on the 2-manifold Hg which may be
thought of as standard. This set of curves is called the Heegaard diagram
of the manifold M3. In order that the curves CJ, C2,.··, cg , ~,ci, ... ,g~
on Hg form the Heegaard diagram of a certain manifold M3 it is neces- Heegaard diagram
sary and sufficient that there hold the following conditions: 1) the curves of a 3-manifold
Ci do not intersect, and after one makes a cut along them one obtains
a connected surface, 2) the same for the curves c:.Referred to as the
Heegaard diagram is sometimes a set of curves {s} on the boundary
Kg (or a set of curves {Ci} on the boundary K;). The corresponding
3-manifold is obtained by glueing 9 handles of index 2 (thick discs) to
Kg along the curves {cn and one handle of index 3 (a ball).

2.5.4 Spines

The classical concept of a spine (not to be confused with "spline") is Spine of a 3-manifold
a convenient means for defining 3-manifolds; it is based on the idea
of reducing the study of 3-manifolds to the study of two-dimensional
polyhedra. See, for example [44].
Coding 3-manifolds by spines is favourable from the viewpoint of a Coding 3-manifo/ds
succesive enumeration of manifolds with their increasing "complexity" by spines
(for details see below). The new theory of almost special spines and its
applications to the problems of three-dimensional topology (including
computer geometry) has been thoroughly developed by S.V. Matveev.
Some of his results admitting a particularly obvious interpretation are
presented in this subsection.

Definition 2. A polyhedron P c M is called the spine of an n-manifold


M with boundary if M -P ~ 8Mx(O, 1). The polyhedron P is called the
spine ofa closed n-manifold if P is the spine of the manifold M - Int Dn ,
where IntDn is an open n-ball in M.

An example of the spine of a manifold is depicted in Fig. 2.5.8. In


Fig. 2.5.9 one can see that any tree (a graph without cycles) in a ball or in
a sphere of any dimension is the spine of the ball or the sphere. A circle Fig. 2.5.8
166 2. Low-Dimensional Manifolds

is a natural spine of a ring 8 1 x I and of a Mobius strip. For any M, the


manifold M - (8M), where O(8M) is a "collar" of the boundary, is the
spine of M.
It is hardly relevant to consider here all possible spines ofa man-
Fig. 2.5.9 ifold. It is convenient to restrict the consideration to the so-called non-
collapsible (or incompressible, closed, non-decreasing, etc.) spines. We
shall recall the defmition of a collapse. Let K be a simplicial complex
(e.g. a triangulated manifold); (In its basic open simplex, i.e. a simplex
which is not a boundary of any simplex from K, and let on-I be its free
boundary, i.e. such that is not the boundary of any other simplex from
K. Then the operation of discarding simplexes (J and 8 from the complex
K is called an elementary collapse (contraction). Figure 2.5.10 presents
collapses of dimensions 1,2 and 3. Collapse is a sequence of elementary
collapses. The collapse of a polyhedron is the collapse of some of its
triangulations. In this case, elementary collapses can be enlarged by in-
cluding in their number the discarding of n-cells together with their free
(n - I)-boundaries (Fig. 2.5.11).
The concept of a collapse appears to be very closely related to the

g. . . . .
concept of a spine.

Fig. 2.5.10 Theorem 2. If P is the spine of a manifold M and if P collapses on Q,


then Q is the spine of the manifold M.

.- j
~
."... The proof of this theorem consists in the fact that the structure of
--~~ the direct product M - P ~ 8M x (0,1) can be changed locally (in the
...........
neighbourhood of some given simplexes) so as to obtain the structure of
Fig.2.5.11 the direct product by M - Q. It is sufficient to verify the possibility of
such a change for a standard simplex in a Euclidean space (Fig. 2.5.12)
because all the simplexes in a manifold are positioned standardly up to
homeomorphism.
A spine is called non-collapsible if it cannot be collapsed into a
smaller spine. From among the above examples of spines it is only the
circle (the spine of a ring or of a Mobius strip) and point (the spine
of a ball or of a sphere) that are non-collapsible for the reason of "the
absence of the beginning", i.e. free boundaries from which the collapse
could start. It can be readily proved that a point is the only non-collapsible
spine of a two-dimensional sphere (a two-dimensional disc). Indeed, a
Fig. 2.5.12
2.5 Visual Properties of Three-Dimensional Manifolds 167

non-collapsible spine of a 2-disc cannot contain 2-simplexes since in that


case there would exist a 2-simplex with a free boundary. Therefore it must
be a polyhedron of dimension not greater than unity and, moreover, a
homotopically contracible polyhedron, i.e. a tree. The only tree without
free vertices is a point. The first part of this consideration is valid for
any dimension.

Theorem 3. Any spine of an n-manifold collapses to become a polyhe-


dron of dimension not greater than n - 1. In particular, any n-manifold
has a spine of dimension :::; n - 1.

The second part of the consideration showing that all one-dimensional


spines of a disc are collapsible into a point is invalid already in three
dimensions: there exist examples of non-collapsible two-dimensinsional
spines of the sphere 8 3• One of such examples is the so-called "Bing
house" with two rooms depicted in Fig.2.5.13 and familar to us from
Chap. 1. The collapse of a b~l onto this manifold is like this. First, along
the tube we penetrate through the upper disc inside the lower room and
exhaust it up to the membrane between the tube and the wall. Then, along
the tube we penetrate through the lower disc inside the upper room and
exhaust it up to the membrane between the upper tube and the wall.
Another example of a non-collapsible tow-dimensional spine of a
ball is a "fool's cap" drawn in Fig.2.5.14. It is obtained from a trian-
gle through identification of its sides. Neither the "Bing house" nor the
"fool's cap" is collapsible for the reason ofthe "absence ofthe beginning"
of the collapse. These examples show that one and the same manifold
may have several different spines. Another ambiguity of this problem is
that distinct manifolds may have homeomorphic (equal) spines. In other
Fig.2.5.14
words, the polyhedron may get thickened up to the manifold in several
ways if at least one thickening does exist. Figure 2.5.15 gives examples
of two thickenings of a circle with diameter up to a manifold of di- :'0. . .-:-.0
I
I
I
I
0

• • •:

I
I
mension two. This is a "genuine buckle", i.e. a disc with two holes and I
I
I
I
I
I
I .: I
a "piratic buckle", i.e. a torus with a hole. Figure 2.5.16 gives similar I

t
I

:
I

J
examples of dimension 3: a full torus with two discarded balls and a ~. . . . . . .I • • • • •"

manifold 8 2 x 8 1 with two discarded balls have one and the same two- Fig. 2.5.15
dimensional spine, namely, the surface of a torus with two discs glued
along meridianal curves.

-Sexiis Fig.2.5.16
168 2. Low-Dimensional Manifolds

2.S.S Special Spines

We introduce the concept of a special spine to eliminate the second


ambiguity - a possible existence of several different thicknesses.

Definition 3. A two-dimensional polyhedron P is called special if the


following conditions are satisfied:
1) Each of its points has a closed neighbourhood homeomorphic
to a cone over a circle with zero,two or three radii (Fig. 2.5.17).
2) Each connected component of nonsingular points (i.e. points
with a neighbourhood of the first type) is a 2-disc.
Fig. 2.5.17
The set of singular points of a special polyhedron is a regular graph
of degree 4, i.e. a graph from each of whose vertices there go exactly 4
edges. Note that singularities of the above-mentioned type are observed in
soap films (the Plateau principles). For details see the book by Fomenko
[29].
The spine of a 3-manifold is called special if it is a special poly-
hedron. Clearly, each special spine is non-collapsible. An example of
the special spine of a 3-sphere is the "Bing house". It has exactly two
vertices (points of the third type), three 2-components (connected com-
ponents of nonsingular points), and the set of its singular points is of the
Fig. 2.5.18 form shown in Fig. 2.5.18.

Theorem 4. Any 3-manifolds has a special spine.

The idea of the proof is as follows. We divide a manifold Minto


handles - one handle of index 0 for each vertex, one handle of index 1 for
each edge and one handle of index 2 for each 2-simplex of an arbitrary
triangulation. Handles of index 3 correspond to 3-simplexes (Fig. 2.5.19).
The union Po of the boundaries of these handles is a special polyhedron
(this is directly seen). We denote the manifold obtained by discarding
an open ball from the interior of each handle by Mo. Then, clearly, Po
is a special spine of the manifold Mo. From this one we can obtain the
Fig. 2.5.19 spine P of the manifold M by puncturing its 2-components and thus
uniting the balls removed into one ball. If the boundary of the manifold
2.5 Visual Properties of Three-Dimensional Manifolds 169

M is non-empty, it must be punctured in one plane. The spine P will


have free edges appeared at the boundaries of the pierced holes, and
therefore it admits collapse and does not, generally, lead to a special
spine. But if puncturing is made carefully lest free edges arise, then we
obtain a special spine. Such a careful puncturing using so-called arches
is presented in Fig. 2.5.20.
The same consideration shows that any 3-manifold has an infmite
number of various special spines.
The advantage of special spines is the possibility of an unambiguous
restoration of a 3-manifold from its special spine. This means that a
special spine carries the whole information about the topological structure
of the manifold.
Fig. 2.5.20

Theorem 5. (see Ref. [44]). If two 3-manifolds have homeomorphic spe- 3-manifolds with
cial spines, they are homeomorphic. homeomorphic special
spines are homeomorphic
The idea of the proof is the following. Since a circle with three
radii is embedded into a two-dimensional sphere in a unique (up to a
homeomorphism) way, the neighbourhoods of the vertices of the special
polyhedron get thickened up to balls in a unique way. This thickening is
uniquely continued up the thickening of the edges of the special polyhe- Special polyhedron
dron. It remains to prove that the thickening of the neighbourhood of the
boundary of each 2-component, which is observed in this case, i.e. the
direct product of this neighbourhood by a segment, is continued up to the
thickening (i.e. the direct product by a segment) of the 2-component itself
in a unique way. This easily follows from the fact that all 2-components
are cells - this is why the condition was imposed.
Similar consideration show that the neighbourhood of the set of
singular points of any special polyhedron (not necessarily the spine of
a 3-manifold) thickens up to the 3-manifold. The continuation of this
thickening up to the thickening of 2-components may be hampered by a
nontrivail character of the fibre bundle of segments appearing near the
boundary. Such an obstacle is really encountered: let a special polyhedron
be obtained by glueing a disc to a projective plane along a projective
straight line. Then it is not embedded into any 3-manifold just for the
reason indicated above. Thus, not all special polyhedra are spines of
3-manifolds, i.e. get thickened. But if a special polyhedron does get Filtration in the set
thickened, it does it in a unique way! of all 3-manifolds
170 2. Low-Dimensional Manifolds

Matveev's complexity 2.5.6 Filtration of 3-Manifolds with Respect to Matveev's


Complexity

Problem of enumerating Theorem 5 implies that the problem of enumerating 3-manifolds reduces
of all 3-manifolds to the problem of enumerating special thickened polyhedra.
An important fact is that for each number k there exists only a finite
number of various 3-manifolds having special spines with k vertices. All
Special thickened polyhedra special polyhedra with k vertices can be enumerated as follows. First we
should construct all regular graphs of degree 4 with k vertices. This is
not difficult to do. Then for each such graph r we should run through all
special polyhedra whose set of singular points is homeomorphic to the
graph r. For this purpose we should do this. We embed the graph r into
"Detail" ]R3 and replace each of its vertices by a so-called "detail" homeomorphic
to the cone above the circle with three radii (Fig. 2.5.21).

IAL
X~ Fig. 2.5.21

Triod Neighbourhoods of triple points of the boundary of a detail - triods


(they are distinguished in Fig. 2.5.21) must correspond to the edges of
the graph which come from a given vertex. For each of the two triodes
corresponding to one and the same edge of the graph r we choose a
homeomorphism of one of them onto the other and glue the details via
2-
chosen homeomorphisms. The boundary of the polyhedron Po obtained
(42.3)
'23 (1 2. 3) (i.e. the set of points of free edges of its arbitrary triangulation) consists
"'ts 2. 1 3 of several circles. Glueing up these circles by 2-discs, we obtain a special
( i 2. 3) ~ (1
"'1S 3) polyhedron the set of whose singular points coincides with the graph r.
2.
13! II o 1 3 2.
(13 2. It3)
~
<\I
( 1 1 3)
3 2. 1 It is clear that any special polyhedron with the graph r taken as the
Fig. 2.5.22
set of singular points can be obtained in this way. Let us analyze the
arbitrariness admitted in the construction. It consists in the choice of
2.5 Visual Properties of Three-Dimensional Manifolds 171

homeomorphisms through which triodes are glued together. For each


two triodes there exist six ways of their identification of which three are
realizable inside the space (we call them even) and three are unrealizable
in R3 (we call them odd). See Fig. 2.5.22.
Thus, there exists exactly 62k possibilites, where k is the number
of vertices, 2k is the number of edges. On running through all these
possibilities, we obtain all special polyhedra corresponding to the graph Fig. 2.5.23
r. The problem of singling out the polyhedra that exhibit thickening
encounters no difficulties: a polyhedron P can be thickened if and only
if in the detour of each circle on the boundary of the polyhedron Po odd
glueings are encountered an even number of times. It can be shown that
there exist not more than 2k- 1 • 32k special polyhedra corresponding to
the graph r, which can be thickened up to an orientable 3-manifold.
We should note the connection between the construction described Fig. 2.5.24
and the Heegaard diagrams (see the end of Sect. 2.5.3). Suppose a closed
orientable 3-manifold M has a special spine P obtained by glueing discs

A·:·:·
to a polyhedron Po glued of details. We thicken all details as shown Polyhedron glued
in Fig.2.5.23. A topologically thickened detail is a 3-ball with two dis- of details
tinguished discs on its boundary, each two of which are joined by a
distinguished arc. Glueing the thickened details gives a solid pretzel. . .
The distinguished arcs will be so glued as to form closed curves on the
0.
........... '.
;... )
pretzel surface which constitute the Heegaard diagram of the manifold
M. So, any closed orientable manifold M has a Heegaard diagram of
a very special form, glued of standard pieces (thickened details). For
any number k there exists only a finite number of special diagrams of
genus k, which fact differs them advantageously from usual diagrams
(the number of the latter is infinite).
The fact that there is a finite number of manifolds having special Fig. 2.5.25
spines with ~. k vertices leads to a natural idea to take as the complexity Complexity of
of a 3-manifold M the number of vertices in its minimal (in the number a 3-manifold
of vertices) special spine. This characteristics is convenient in that there
exists only a finite number of3-manifolds of complexity k. But it has its
shortcomings too: first, the complexity of a sphere appears to be equal
to 1, whereas that of a manifold £3,1 appears to be equal to zero. Figure
2.5.24 presents the neighbourhood of a singular graph of the minimal
special spine of a sphere, and in Fig. 2.5.25 we can see the neighbourhood
of a singular graph of the minimal spine of a lens £3,1, which is a triangle
with identified compatibly oriented edges (we shall call it the "Mobius
cap").
172 2. Low-Dimensional Manifolds

Second, the complexity of a connected sum of two manifolds is not


generally equal to the sum of their complexities. These two shortcomings
have one and the same root: a non-collapsible subpolyhedron of a special
polyhedron is not necessarily special. Let us introduce the concept of an
almost special polyhedron free of this shortcoming.

Almost special polyhedron Definition 4. A non-collapsible polyhedron P is called almost special


if each of its points has a neighbourhood homeomorphic either to a
polyhedron of dimension ~ 1 or to a union of such (perhaps empty) a
polyhedron with the cone above the circle (the latter may have two or
three radii).

Points, in any neighbourhood of which there enters a cone above a


Vertices of almost circle with three radii, are called vertices of an almost special polyhe-
special polyhedron dron. Each almost special polyhedron is a union of a one-dimensional
polyhedron and a two-dimensional polyhedron which differs from a spe-
cial one in that it may have 2-components non-homeomorphic to a disc.

~~~
~~Y
An example of an almost special polyhedron is given in Fig. 2.5.26. The
polyhedron porperty of being almost special is preserved when we pass
over to the non-collapsible polyhedron. This follows readily from the
definition.
Fig. 2.5.26
Complexity of Definition 5. (S.V. Matveev). The complexity of a 3-manifold is equal
a 3-manifold to k if it has an almost special spine with k vertices and has not a single
special spine with a smaller number of vertices.

Manifolds 8 3 ,82 X 8 1, m>3 and L3,1 have zero complexity. Their


Minimal almost minimal almost special spines are respectively a point, a bouquet of a
special spines sphere and a circle, a projective plane and a Mobius cup. The connected
sums of these manifolds also appear to be of complexity O. It turns out
that there exist no other closed orientable manifolds of complexity O.

Additive property Theorem 6. (S.V. Matveev). Let the complexity of a closed orientable
of the complexity 3-manifold M be equal to k. Then M = Ui Mi, where each Mi is homeo-
morphic to 82 x 8 1, m>3 or has a special spine Pi the number of whose
vertices coincides with the complexity ki of the manifold Mi. Given this,
there holds the equality k =Ei ki.
2.5 Visual Properties of Three-Dimensional Manifolds 173

The idea of the proof is the following. If the minimal almost special Idea of the proof
spine P of the manifold M contains points with one-dimensional neigh-
bourhoods, it contains the principal edge, i.e. an edge not adjoint by the
two-dimensional part of the polyhedron P. When this edge is cut, the
spine P either falls into two parts, which corresponds to the manifold
decomposition into a connected sum, or does not fall into parts at all,
which corresponds to singling out the summand 8 2 x 8 1. If some of the
2-components of the polyhedron P contains a simple closed curve not
restricting a disc, then the polyhedron P should be cut along this curve,
and one of the circles resulting from the cut should be glued up by a
disc. The manifold will then fall into a connected sum of two manifolds
with simpler spines. If the spine of P contains no principal edges and
if all simple closed curves in its 2-components restrict discs, then this
polyhedron is either special or homeomorphic to ]R2, which corresponds
to the case of projective space. The proof of the inequality 2:i ki ~ k is
obvious since if we join the minimal almost spines of the manifolds Mi
with arcs, we obtain the almost special spine of the manifold M. The
inverse inequality 2:i ki ::; k is proved using the Haken theory of normal Haken theory
surfaces [45], and the proof is rather cumbersome. of normal surfaces
This theorem implies the above-mentioned result concerning ori-
entable manifolds of complexity 0 since the only special spine with 0 3-manifolds
vertices is a Mobius cap. From the theorem it also follows that up to con- of complexity 0
nected sums with manifolds 8 2 x 8 1, ]Rp3, L 3,1 there exists only a finite
number of manifolds of complexity k. In other words, if we restrict our-
selves to closed orientable manifolds containing no non-splitting spheres, Only a finite number
projective planes and Mobius caps, then there exists only a finite number of irreducible manifolds
of such manifolds of complexity k, and the minimal almost special spine with a fixed complexity k
of such a manifold is special.

2.5.7 Simplification of Special Spines

As mentioned above, a 3-manifold may have an infinite number of vari-


ous special spines. How are different special spines of one and the same
3-manifold connected? How shall we describe the set of all special spines Set of all special spines
of this manifold and how is the minimal one to be distinguished among
them? We shall describe the technique, called re-laying of2-component, Re-Iaying of 2-components
174 2. Low-Dimensional Manifolds

Simplification of which enables one special spine to be transformed into other special
special spines spines. For simplicity we shall consider only closed 3-manifolds. Let P
be a special spine of a closed 3-manifold M and let D c M be a disc
in M having no common interior points with P, such that the curve aD
lies in P and is in general position to its singular graph r. This means
that the curve aD does not pass through the vertices of the graph rand
crosses its edges transversally. The disc D splits an open ball M - P
into two balls Vi and Y2. An important and easily verified (proceeding
from the structure of singular points of the polyhedron P) fact is that
the intersection K of closures of the balls Vi and Y2, i.e. their common
boundary is a closed 2-manifold composed entirely of 2-components of
the polyhedron P. The polyhedron MUD is a speGial spine of the man-
ifold M with two removed balls. If we puncture the disc D, we go back
to the spine P of the manifold M with one removed ball, i.e. simply to
a spine of the manifold M. But if we puncture one of the 2-components
on the surface K, then after collapse we obtain another special (or al-
most special) spine PI of the manifold M. Figure 2.5.27 presents two
particularly simple re-Iayings TI and T2 made in the neighbourhood of
a vertex or an edge of the polyhedron P. Inverse transformations TI- I
and T2- 1 are also re-Iayings.
It turns out that even these very simple re-Iayings suffice to transform
one special spine of a 3-manifold into another.

Theorem 7. (S.V. Matveev). Two special polyhedra are special spines


of one and the same 3-manifold if and only if they can be joined by the
chain of transformations T~I , T2±1.

The proof of this theorem is based on the fact that any two tri-
angulations of a 3-manifold have a common subdivion. Note that the
complexity (the number of vertices) of a special polyhedron increases
by 2 under transformation TI and by 1 under transformation T2. Let us
Fig. 2.5.27
analyze the change in the complexity of a special polyhedron under an
arbitrary re-Iaying. Let a curve aD intersect the graph r at k points
and let the boundary of the punctured 2-component a of the surface K
contain m vertices. After the disc D is glued to the polyhedron P, the
number of its vertices increases by k, and after the 2-component is punc-
tured and the polyhedron collapses, the number of its vertices decreases
not less than by m. Therefore, the complexity p,(PI) of the polyhedron
2.5 Visual Properties of Three-Dimensional Manifolds 175

obtained as a result of re-Iaying of the special polyhedron PJ does not


exceed J.L(P) + k - m, where J.L(P) is the complexity of the polyhedron
P. We shall call the re-Iaying simplifYing if J.L(PJ) < J.L(P) and balanced Simplifying re-Iaying
if J.L(PJ) = J.L(P). The number k will be called the degree of re-Iaying.
Some experience is dealing with spines shows that re-Iaying of degree Degree of re-Iaying
~ 6 are of particular importance.

Hypothesis. Any special spine of a 3-manifold is reduced to any mini-


mal (in the sense of the number of vertices) one using simplifYing and
balanced re-layings of degree ~ 6.

Note that a positive solution of this hypothesis would imply the pos-
sibility of algorithmic recognition of homeomorphic 3-manifolds. This
follows from the fact that the number of special polyhedra of fixed com-
plexity is finite as well as from the fact that re-Iayings of fixed degree
can be applied to a given special polyhedron also in a finite number of
ways. Let PJ and P2 be special spines of manifolds MJ and M2. We shall
apply to them simplifying and balanced re-Iayings as long as possible. Balanced re-Iaying
If the two finite sets of special spines which we obtained intersect, then Hypothesis about
MJ and M2 are homeomorphic and if not, then MJ and M2 are distinct. algorithmic recognition of
The validity of the hypothesis can be grounded as follows. homeomorphic 3-manifolds

1) The hypothesis is valid for all manifolds of complexity ~ 6. Validity of this


Practically, re-Iayings of degree ~ 4 are sufficient. conjecture for special
2) The hypothesis is valid for the series of special spines of a sphere types of a 3-manifolds
which are constructed in a canonical way using the Heegaard diagram
of a sphere simplifyied by the Volodin-Fomenko cut-vertex and "wave" Cut vertex
method [29]. and wave method
3) The hypothesis is valid for special spines of a sphere obtained
using the construction of "combinatorially disassembled" (i.e. remain-
ing after removal of 3-simplexes collapsing to a point) triangulations
described in the proof of Theorem 6.
We shall point out the simplest and therefore most easily applicable
in practice particular cases of simplifying re-Iayings, i.e. simplification
by the counterun and by the short boundary curve. Recall that any spe- Boundary curves
cial polyhedron is obtained from a certain regular graph of degree 4 by of 2-components
glueing several 2-discs via some maps of their boundary circles. The
curves in the graph corresponding to these maps will be called boundary
176 2. Low-Dimensional Manifolds

Counterrun curves of 2-components; We say that a boundary curve has a counterrun


if it passes on one of the edges of the graph twice in opposite directions.
Short boundary curve We say that a boundary curve is short if it passes through not more than
three vertices of the graph, once through each.

Theorem 8. If a special spine of a closed 3-manifold has a boundary


curve with counterrun or a short boundary curve, then it is not minimal.

p Proof. Suppose in a special spine P of a closed 3-manifold M there


exist a boundary curve with counterrun. Then in M there exists a disc D
which has no common interior points with the polyhedron P and whose
Fig. 2.5.28 boundary contains exactly one singular point and lies in a 2-component
Re-/aying procedure with counterrun on the boundary (Fig. 2.5.28). The re-Iaying procedure
consisting in adding the disc D to the spine P and puncturing one of the
side walls (as shown in the figure) is simplifying. Indeed, in puncturing
and collapse there vanish at least the vertex added and the singular edges,
starting at this vertex, together with other vertices (the endpoints of these
edges).

Suppose now that in the special spine P of the closed 3-manifold


M there exists a 2-component a: with a short boundary curve passing
hole through k ::; 3 different vertices. If inside the manifold we glue to P a
Fig. 2.5.29
disc parallel to the component a: and puncture the side wall of the cylinder
obtained, then after a collapse we obtain an almost special polyhedron
with a smaller number of vertices. Indeed, k vertices are added and at
least 2 vanish for k = 1 and 4 for k = 2,3 (Figs, 2.5,29, 30).

Note that one can easily realize the indicated simplification technique
and construct, in fact, an almost special spine with a smaller number of
Bing house vertices. The Bing house has both a short boundary curve and a counter-
run. Using any of these procedures, one can simplify the Bing house to
a point. It is noteworthy that in the case of a boundary curve of length
4, such a re-Iaying is balanced.
2.5 Visual Properties of Three-Dimensional Manifolds 177

2.5.8 The Use of Computers in Three-Dimensional Topology. 1<d


Enumeration of Manifolds in Increasing Order of Complexity

The results of subsections 2.5.6 and 2.5.7 allow enumeration of all closed
orientable (containing no non-splitting spheres, projective planes and
Mobius caps) manifolds using computer. Such an enumeration of mani-
folds of complexity :::; 5 was made in 1972 by S.V. Matveev and V.V.
Savvateev and then repeated in an extended version (including some disk hole
manifolds of complexity 6) by S.V. Matveev in 1984 [46]. In 1986, S.V.
Fig. 2.5.30
Matveev enumerated all manifolds of complexity 6 [47]. Recently, in
1987, Matveev and Fomenko [48] used computer for analysis one of the Computers in 3-topology
problems of three-dimensional topology, namely, for estimation the least
volume of a compact hyperbolic manifold. Restriction to complexity 6
is caused both by the lack of machine time (calculation of each graph
with 6 vertices takes up 5 to 8 hours depending on its type) and by
difficulties of manual identification of manifolds yielded by computer.
Regular graphs of degree 4 with a given number of vertices are to be
enumerated manually (for the first time this was done in 1972) or using Enumeration of
computer (1984). Then each graph is set into computer which enumer- a 3-manifolds
ates all possible glueings to toroides (see Sect. 2.5.6) and subjects the Glueings of toroides
boundary curves of the special polyhedron obtained to the check:
1) there exists a boundary curve oflength :::; 3;
2) there exists a counterrun;
3) the glueing of thickened details gives Heegaard diagrams of the Glueing of
orientable manifold. thickened details

If the answer to one of the first two questions is affirmative and


to the third question is negative, the given choice of toroide glueing is
rejected and a new version is set for verification. Otherwise the com- Results of
puter calculates the one-dimensional homoloy group of the manifold and computer experiment
conveys the result, along with the triode glueing system, on a printer.
The results of computer calculations are listed in Tables 1-9. The tables
include all irreducible manifolds of complexity :::; 5 and those manifolds All irreducible 3-manifolds
of complexity:::; 6 which were modelled on the one-dimenisonal octa- of complexity ::; 5
178 2. Low-Dimensional Manifolds

hedron frame, on a long figure-of-eight and on a closed chain. Near the


picture of the graph there are closed orientable 3-manifolds modelled on
this graph and satisfying the check on the basis of criteria I, 2, 3. Here
Lp,q denotes a lens space with parameters p, q and 83/G is a manifold
Quotient spaces with a fmite fundamental group whose universal covering space is the
of a 3-sphere sphere 8 3 , or in other words the quotion space of a sphere by a certain
List of the groups free action of the group G. These groups are as follows:
with a free action
I) Q4n = {x,y;x 2 = (xyi = yn}, where n ~ 2;
2) P24,P48,P120 = {x,y;x 2 = (xy)3 = yn;x4 = I}, where n =
3,4,5 respectively.
3) D2k(2n+l) = {X,y;x 2k = l,y2n+l = l,xyx- 1 = y-l}, where
k ~ 3,n ~ 1.
4) P~2 = {x,y,z;x 2=(xy)2=y2,zyz-l=xy,zxz- 1=y,z9=1}.
In all the cases the subscript indicates the order of the group.
5) The direct products of the groups listed above by the cyclic
group ate of mutually simple order.

Description of a These groups are known to be able to act freely and linearly
corresponding 3-manifolds on the sphere 8 3 (J. Milnor). In particular, the groups Q4n, P24, P48
and P120 are subgroups of the groups 8 3 (of unit length quatemions).
In the two-sheeted covering 8 3 ~ Jltll3 = SO(3) these subgroups
cover twice the orientation-preserving symmetry groups of a regular
n-gon prism (dihedron), tetrahedron, octahedron and dodecahedron re-
spectively. Denoted by T2 x 1/ (~ ~) is the 3-manifold obtained
by glueing the bases of the direct product of a torus by a segment
via the homeomorphism specified by the indicated matrix. Denoted by

(~: ~~) \ T2 XI/ (~: ~~) is the manifold obtained by identi-


fication of points of one base of the direct product T2 x I by involution
with the matrix

and the second with the matrix ( a2 b2 )


C2 d2
2.5 Visual Properties of Three-Dimensional Manifolds 179

For each of the manifolds underscored in Tables 4-9 there is a drawing


of the neighbourhood of the singular graph of its minimal special spine. List of singular graphs
The choice of these manifolds was made by the first appearance principle of a special spines
in Tables 1-3. The exception was made for lens spaces whose minimal
special spines were modelled on long figures-of-eight and look uniformly.
We therefore restrict ourselves to one arbitrarily chosen example of each
complexity.
We provide the list of all irreducible (Le. containing no nontriv- List of all irreducible
ial spheres) closed orientable 3-manifolds of complexity ~ 5 with a closed orientable 3-manifolds
noncommutative fundamental group. For manifolds with a commutative of complexity ::; 5
fundamental group of complexity ~ 5, the Heegaard genus is ~ 1. So,

8 3/Q8, 8 3/Q12, 8 3/P24, 83/ D24, 83/Q8 X Z3, 83/QI6,


8 3/ P120, 8 3/ P48, 8 3/ P42' 8 3/ D40, 8 3/Q16 X Z3, 8 3/Q20 X Z3,
8 3/Q8 X Zs,83/Q12 X Zs,83/D48,8 3/P24 X Zs,8 3/Q20 .

It= 0

Table 1
180 2. Low-Dimensional Manifolds

n=S'

Table 2
2.5 Visual Properties of Three-Dimensional Manifolds 181

i L1.l,2..'
s,1'
--
L19,3,Lzo,3,L2f,J,.,L2.3,J,. , L2.4-,5" ,
.

L2.4;11'L2.1,S' L2.8,S' L2.9, g' L30, '/, L31, ,/', L 3 (,ff '

L32: 1, L33.r,L33,10,L31t;9,L3S,S' L 36,11,i.J 3 7;8 J

1-.3'1, 10'L39, 11t' L 39 ,16} L t,.0,11, L 1If,1f,L 4f, 16' L 1,.3.12. ,


L 4-Ir,13 ,L IrS, 19 ,L Irs d2 ,LI,.6,11, "'41,13 ,L 4-9 (8 J
L so, 19 , L,j~2f' ,

o.
Table 4
182 2. Low-Dimensional Manifolds

Table 7
2.5 Visual Properties of Three-Dimensional Manifolds 183

Table 8

Table 9
184 2. Low·Dimensional Manifolds

Matveev's complexity 2.5.9 Matveev's Complexity of 3-Manifolds and Simplex Glueings


of 3-manifolds
As mentioned above, if we take several triangles, divide their edges into
Connection with pairs and glue the edges of each pair, we obtain a closed 2-manifold.
simplex glueings Doing this, we obtain any closed 2-manifold. Let us see what we have
in three dimensions.

4 -~.. )
We take several oriented tetrahedra, divide their faces into pairs and

(B. [J .,
~.- glue faces of each pair through one of the three possible orientation

®
.... ......
reversing homeomorphisms. This will give us a topological space (a cell
complex, a polyhedron) X. We shall attempt to prove that X is a closed
• • I 1
~
\. ... - J.!
: I t -., ; '1.5
~
orientable 3-manifold. If a point x E X corresponds to an interior point
Fig. 2.5.31 of one of the tetrahedra, it has in X a ball neighbourhood since there was
such a ball neighbourhood in the tetrahedron. If the point x corresponds to
a pair of points inside two faces of tetrahedra, then its ball neighbourhood
is glued of two half-ball neighbourhoods (Fig. 2.5.31). Let the point x
corresponds to points ai, a2, ... , an on the edges. Each point on the edge
has in its tetrahedron a neighbourhood of a ''watermelon slice" type. Of
them one glues the ball neighbourhood of the point x (Fig. 2.5.32).
Let the point x be obtained by glueing vertices. Each vertex has in
the tetrahedron containing it a cone-like neighbourhood over the triangle.
Fig. 2.5.32 When glued, such cone-like neighbourhoods form the neighbourhood of
the point x, homeomorphic to the cone over an ,orientable closed surface.
Cone·like singularities Thus, the space X is a 3-manifold with cone-like singularities over a
surface. In case all the surfaces are 2-spheres, X is a closed 3-manifold.
There is a simple way to establish this fact.

Theorem 9. The space X obtained by glueing faces of tetrahedra is a


Euler characteristic 3-manifold if and only if its Euler characteristic is equal to zero.

The necessity of the equality x(X) = 0 follows from the fact that
the Euler characteristic of any closed 3-rnanifold is equal to zero, which
can in turn be deduced from the existence of Heegaard splitting. Indeed,
if Kn and K~ are solid pretzels of Heegard splitting of genus n, then
X(X) = X(Kn) + X(K~) - X(Kn n K~) = 1 - n + 1 - n - (2 - n) = 0,
where Kn n K~ is their common surface.
Sufficiency of the equality X(X) = 0 is deduced from the following
considerations. Let X have cone-like singularities at points a], a2, ... ,an
2.5 Visual Properties of Three-Dimensional Manifolds 185

over surfaces FI , F2, ... ,Fn. Removal of the open cone-like neighbour-
hood of the point ai results in a decrease of the Euler characteristic by the
number 1- x(Fi) since this procedure removes one vertex (the point ai)
and as many edges; triangles and tetrahedra as there are vertices, edges
and triangles on the surface Fi. On removing open conic neighbourhoods Conic neighbourhoods
of all singular points, one obtains a three-manifold Y with boundary and of a singular points
with the characteristic X(Y) = X(X) - Ei=I(l - x(Fi». On the other
hand, by means of doubling the manifold Y one easily derives the for-
mula X(Y) = 1/2 Ei=1 X(Fi). Comparison of these two equalities gives
the new one X(X) = 1/2 Ei=1 (2 - X(Fi». Since no characteristic of a
surface exceeds 2, in the case X(X) = 0 we have X(Fi) = 2, ie. all Fi
are spheres. It turns out that there exists a close relation between the Relation between
complexity of a 3-manifold and the number of simplexes necessary for complexity and the
glueing this manifold. number of simplexes

Theorem 10. The complexity of an imducible closed orlentable 3-


manifold does not exceed the number k f 0 if and only if it can be
obtained by glueing ~ k tetrahedra.

The proof of this theorem is based on the following remark. In a


tetrahedron we take a cone with the vertex in the centre of the tetra-
hedron over the union of short segments of all medians of all its faces
(Fig. 2.5.33). Cut the angles of the tetrahedron, ie. remove the conic
neighbourhoods of the vertices.
Remove also the neighbourhoods of the type of the direct product
of a circular sector by a segment in the remaining part of the edges
(Fig. 2.5.33). As can be readily seen, this gives a thickened detail with
a standard detail (see Sect. 2.5.6) inside. Suppose the complexity of the
manifold M does not exceed k. Then M has an almost special spine
with k vertices. Since M is irreducible, this spine is special by Theorem
6. The neighbourhood in M of its singular graph is obtained by glueing ~
thickened details ie. tetrahedra with removed neighbourhoods of the I.:::::Y
vertices and edges. The addition to this neighbourhood of those of 2-
components is equivalent to the addition to the tetrahedra of the removed
neighbourhoods of edges and yields M without a ball. The addition of
this ball is equivalent to the addition to the tetrahedra of the cut-out angles
and leads to the unknown glueing of the manifold M of k tetrahedra.

~~ '2! ~
Fig.2.5.33
186 2. Low-Dimensional Manifolds

Inversely, if M is obtained by glueing k tetrahedra then the cones over


their meridians are glued to form a special spine with k vertices of the
manifold M with several balls removed. Their number coincides with
that of the points obtained by glueing the vertices of the tetrahedra. The
union of these balls, obtained by means of puncturing the walls between
them, and the collapse give a spine of the manifold M with ~ k vertices.
We shall show on an example how this result can be applied. What
manifolds can be obtained by glueing faces of one octahedron? In his
well-known preprint "Geometry and Topology of 3-Manifolds" Thurston
Fig.2.5.34b presents a rough estimate of their number: 8505 (i.e. = 8!3 42- 4(4!)-1).
There are practically much less closed manifolds of this type. They can
all be enumerated. Indeed, on octahedron is glued of four tetrahedra
(Fig.2.5.34a), and therefore an irreducible manifold of this type must
have complexity ::; 4. All such manifolds are listed in Table 1. Mani-
folds corresponding to the graph of Fig. 2.5.34b cannot be obtained since
Fig.2.5.34c
tetrahedra are already partially glued in the octahedron by the scheme of
Fig. 2.5.34c which is absent in a long figure-of-eight.
Tetrahedron and For example, gluing each face of a tetrahedron to an opposite face
3-manifolds SI P24 by superposition of parallel transport and rotation through an angle of
60° leads to the manifold S3 / P24. Overturning each upper face onto a
lower one gives a sphere. If at the same time two disjoint upper faces
are rotated through an angle of 120°, we obtain a connected sum of two
copies of the lens L3,1 (Fig. 2.5.35).

Fig. 2.5.35

Page 187, Fig. 2.5.36. Spines of a 3-manifolds can be considered as the


codes of a 3-manifold. Ancient Indian legend about creation of the World
188 2. Low-Dimensional Manifolds

2.5.10 Visual Material

A series of figures (Figs. 2.5.36, 37, 38, 39) illustrates the scheme of
Coding of 3-manifolds coding three-dimensional manifolds by means of spines. For this pur-
by means of spines pose we first cut a manifold (Fig. 2.5.36) and remove from it, say, one
three-dimensional cell (a ball). We obtain a 3-manifold with bound-
ary. In Fig. 2.5.36 the operation is shown" literally" as cutting a three-
dimensional body (volume) with a knife. Then the manifold with bound-
ary starts contracting to its special two-dimensional frame (= spine) de-
scribed in Sect. 2.5 above. Spines can be thought of as piecewise-linear
Flexible cells polyhedra, as shown in Fig. 2.5.37. Here flexible, amorphous cells of the
smooth 3-manifold gradually contract to rectilinear frames and freeze in
this form as the "code of the manifold". Given this, each manifold can be
assignes "complexity", i.e. the minimal number of vertices in the almost
Filtration in the set special spine. As a result, the codes of all 3-manifolds form an infinite
of all 3-manifolds sequence in the increasing order of their complexity. Such an ordering
(filtration) is schematically shown in Fig.2.5.38. The farther we move
along this sequence, the more complicated became the codes of the mani-
folds. Contraction of 3-manifolds to its spine is schematically shown also
in Fig. 2.5.39 where flexible cells of the manifold are shown as bended,
rather amorphous objects and the spine of the manifold as rectilinear (cu-
bic) constructions. All the cells gradually "settle" of the spine, and the
code of the manifold is transformed into a two-dimensional polyhedron.
Extra-mathematical Figure 2.5.39 has also some extra-mathematical content. It can be
pictures considered as an illustration for the novel "Master and Margaret" by the
famous Russian writer Michael Bulgakov. The giant cat Begemot (the
servant of the devil Woland) watchs the crucifixion of Jesus Christ.

Page 189, Fig. 2.5.37. Spines and complexity of a 3-manifolds. Crucifixion


according to medieval Roman legends

Page 190, Fig. 2.5.38. Filtration in the set of all 3-manifolds. Gods of
ancient India

Page 191, Fig. 2.5.39. Retraction of a 3-manifold onto its spine. To M.


Bulgakov's "Master and Margaret": mystical cat Begemot watchs the cru-
cifixion of Jesus
3. Visual Symplectic Topology and Visual
Hamiltonian Mechanics

3.1 Some Concepts of Hamiltonian Geometry

3.1.1 Hamiltonian Systems on Symplectic Manifolds

We shall consider a vector field v on a smooth manifold. Relative


to local coordinates Xl, ... , xn this field can be written as dx i / dt =
vi(xl, ... ,xn) where I ::; i ::; nand vi(xl, ... ,xn) are smooth func-
tions which are components of the field v. Thus, each vector field is Vector fields
interpreted as a system of ordinary differential equations on a manifold. as differential equations
Inversely, each system or ordinary differential equations can be repre-
sented as the vector field on a corresponding manifold. Many physical
laws are written in the form of differential equations, and therefore the
study of topological properties of vector fields provides a good deal
of qualitative information on the behaviour of certain physical systems.
Among the variety of mechanical and physical systems there exists an
important class described using so-called Hamiltonian equations. These Hamiltonian equations
systems can be realized on even-dimensional manifolds only.
A manifold M2n is called symplectic if a skew-symmetric closed Symplectic manifolds
scalar product on this manifold is defined, i.e. in each tangent plane Skew-symmetric
TxM to the manifold M at the point x a nondegenerate skew-symmetric scalar product
bilinear form (a, b) is given (on tangent vectors a and b to the manifold) SymplectiC 2-form
which depends smoothly on the point and which determines the closed as symplectic structure
differential 2-form w, i.e. dw = 0, where d is the exterior differential
operator. This form can be interpreted as a skew-symmetric scalar product
of the vectors a and b, i.e. (a, b) = -(b, a). See [49].
194 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Euclidean space ]R2n Example. The Euclidean space JR.2n referred to Cartesian coordinates
as a symplectic manifold pi , ... ,pn , ql , ... ,qn and endowed with a bilinear form whose matrix
. ( -E0 0
IS E) ,were .matrIx.
h E'IS an n x n umt . It turns out that m.
the neighbourhood of any point on an arbitrary symplectic smooth mani-
fold one can always choose such local coordinates pi, ... ,pn, ql , ... , qn
(already not necessarily locally Euclidean) relative to which the matrix
of the form (a, b) will be written as ( _ ~ ~) independent of the
point (in the given neighbourhood). In other words, if (a, b) is written
in an explicit form '£WijaiQi, where Wij = -Wji and [} = (Wij) is the
skew-symmetric matrix of the form (a, b), then there exists a coordinate
change in the entire neighbourhood of the given point, such that at each
point of this neighbourhood the matrix of the form is (_~ ~) (the
Darboux theorem Darboux theorem). Such coordinates are called symplectic.
Symplectic coordinates
Definition.1. Let v be a smooth vector field (a system of differential equa-
tions) on an even-dimensional symplectic manifold M2n. This system is
Hamiltonian system called Hamiltonian if on the manifold there exists a smooth function H
Energy or (termed the Hamiltonian or the Hamiltonian/unction) such that in local
Hamiltonian function sympIt ' coord't
ec IC ma es pI , ... , pnl
,q , ... , qn(h'h
W IC are known t 't
0 eXIS
always in the neighbourhood of any point of the manifold) this system
Canonical Hamiltonian form v is written in the (so-called Hamiltonian) form:
of differential equation
dpi = oH(p, q) dqi = _ oH(p, q)
dt oqi dt opi

For brevity, df /dt will be sometimes denoted by j. Thus, a Hamil-


tonion system in appropriate coordinates is written as a vector field v =
Physical Hamiltonian (pi, ... ,pn, ql , ... , qn) with components of the form pi = 0H/ oqi , qi =
equations -oH/opi. Many fundamental physical equations have precisely this
form (provided the coordinate system is properly chosen). This accounts
for the interest of both mathematicians and physicists in Hamiltonian
equations.
3.1 Some Concepts of Hamiltonian Geometry 195

Recall that a function I on a manifold is called the integral of the Integrals of the
vector field (the system) v if its derivative along this field (i.e. the func- dynamical system
tion v(f) = 1:. viaf/ 8xi) is identically zero. Equivalently, the function
I is the integral of the field v if and only if it is constant on all inte-
gral trajectories of the system. An integral trajectory of a system is a Integral trajectofy
smooth curve whose velocity vector coincides at each of its points with
the vector of the field v (at this given point).
One can readily see that a Hamiltonian is always the integral of a Function of Hamilton
Hamiltonian field. Indeed, is always the integral

If the function I is the integral of the field v this field is tangent


to the level surface of the function I (see the definition of the inte-
gral). Consequently, each Hamiltonian field is always tangent to the
level surfaces of its Hamilitonian H (the energy integral). To say it dif-
ferently, the hypersurfaces described by the equation H = const (if they
are nonsingular, their dimension is obviously equal to 2n - I) are in-
variant under the field v, i.e. the integral trajectories of the field occupy
these surfaces never to leave them (Fig. 3.1.1). Thus, any Hamiltonian
Fig.3.1.1
field always has at least one integral. If the field is not Hamiltonian
then such a theorem does not certainly hold. Suppose the field v has
two functionally independent integrals H and I. Recall that two func- Functionally
tions are called functionally independent on a manifold if their gradients independent integrals
grad H and grad I are linearly independent almost everywhere (i.e. on
an open set of points which is everywhere dense in the manifold). Then
the field v is tangent both to the surfaces H = const and I = const, Isoenergy surlaces
and therefore it is tangent to the (2n - 2)-dimensional surfaces which
are intersections of the (2n - I)-dimensional surfaces H = const and
1= const (Fig. 3.1.2), and so on. If the system possesses k functionally
independent integrals H = II, h ... ,/k then it preserves all level sur-
faces Ii = const, 1 ::; i ::; k, and all their intersections, and therefore· the
integral trajectories of the system lie on the (2n - k)-dimensional sur-
faces which are common level surfaces of all the integrals II,···, Ik.
These common level surfaces are given by the system of equations
11 = const, ... , !k = const. Consequently, the more integrals the system Fig.3.1.2
196 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

has, the lower its order. This explains the desire of specialists investigat-
ing one or another system of equations to find as many of its independent
integrals as possible.

Involutive integrals 3.1.2 Involutive Integrals and Liouville Tori

The above arguments can be applied to any vector field. In the case
of Hamiltonian fields there is an important specific fact which in some
cases permits, roughly speaking, the order of the system to be lowered
each time not by unity but by two unities, i.e. each time the integral is
so to say "taken for two integrals". This happens when the system has
integrals in involution (involutive integrals). We recall the defmtion of
Poisson bracket the Poisson bracket of two functions. Let I and 9 be smooth functions
of the functions on a symplectic manifold M. We choose local symplectic coordinates
Some dynamical systems pi, ... ,pn, ql , ... ,qn on M2n. The Poisson bracket {f, g} on the func-
have many integrals tions I and 9 is the function given by the formula
which are in involution
n 01 og 01 og
{f, g} = ~ opi oqi - oqi opi

Definition 2. Two functions I and 9 are said to be in involution (to


commute) if their Poisson bracket is identically zero.

In mechanics and physics ther occur situations when a Hamiltonian


system v on a symplectic manifold M2n has n independent integrals
H = II, 12.' :.. ,In which are in involution. It turns out that in this case
one can say rather much about the behaviour of the solutions of the
Liouville system (i.e. its integral trajectories). Such systems are called Liouville
integrable systems = integrable or completely integrable systems. Let us consider a common
completely integrable level surface rn of all n integrals, i.e. the one given by equations II =
differennalequanons CI, .•. ,In = en (where Ci are some constants). Suppose for simplicity
that this surface is connected and compact and suppose the integrals are
independent on this surface. Then this sUrface proves to be diffeomorphic
Liouville theorem to an n-dimensional torus rn, i.e. to the direct product of n circles (the
Liouville tori Liouville theorem). These tori are occasionally referred to as Liouville
tori.
3.1 Some Concepts of Hamiltonian Geometry 197

Next, the initial system leaves these tori invariant, i.e. the integral
trajectories (solutions) of the system lie on the tori and fill them com- Special coordinates
pletely. Furthermore, on a Liouville torus and in its neighbourhood one in the neigbourhood
can always choose coordinates in which the system v will be written of a Liouville torus
as that with constant (!) coefficients. Therefore, in these coordinates the Angle-action variables
integral trajectories will determine the rectilinear winding of the torus. Rectilinear winding =
For example, if the torus is two-dimensional, it can be represented as a almost periodic motion
glued square (Fig. 3.l.3). Then (in appropriate coordinates) the integral
trajectories of the field are depicted as straight lines (Fig. 3.1.3). In this
situation two cases are possible: a) all integral trajectories (straight lines)
are closed on the torus; b) all of them are non-closed i.e. are wound
round the torus infinitely long and without self-intersections. In the first
case, all integral trajectories are homeomorphic to a circle (periodic so- a
lutions of the system). In the second case, all of them are homeomorphic Fig.3.1 .3
to a straight line. In the second case the system is said to determine the
conditionally (or almost) periodic motion (and periodic in the first case). Periodic =
In the model of a torus on a square it can be readily explained when rational solutions
the case "a" and when the case "b" is realized. We consider the angle a Almost periodic =
between the straight line and the x-axis in Fig. 3.1.3. If the tangent of the irrational solutions
angle a is rational, the integral trajectory will sooner of later close up on
the torus. The point is that a straight line on a plane gets sooner or later
to the point with integer-valued coordinates (m, n), where the mutually
simple numbers m and n are such that tan a == min. If the tangent of the
angle a is irrational, the straight line passes by all integer points (m, n)
on the plane (except for its starting point (0,0)), and therefore after
the plane is rolled up into a torus, the trajectory cannot get closed. In
particular, it never returns to the initial point and starts winding infinitely Irrational trajectories
round the torus, this covering being increasingly dense. It can be shown are everywhere dense
that this torus winding is everywhere dense, i.e. if we fix an arbitrary
point on the torus and an arbitrary small number c: > 0, then the moment
is sure to come when our trajectory will pass near the chosen point at
a distance not exceeding the number c:. In other words, in some time
such a trajectory will pass arbitrarily close to any aforegiven point of the
torus. True, it may be a long time before this event happens.
If we now close such an irrational trajectory, we obtain the whole
torus. In this case the Liouville torus is a set of points arbitrarily closely
approximated by points of the given integral trajectory. It is known from
198 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Liouville torus mechanics, that for analytic systems the case described above is the case
as the closure of general position, i.e. typical in the sense that if we choose randomly
of its irrational the initial point of the integral trajectory, send a trajectory from this point,
integral trajectory continue it to infinity and close, we obtain a complete Liouville torus.
The cases when an integral trajectory returns to the initial point (i.e. is
homeomorphic to a circle) are "exceedingly rare" and correspond to so-
called resonances. In systems of general form, such resonance points fill
up a set of measure zero, i.e. a very "thin" set.
Thus, the behaviour of integral trajectories of Liouville-integrable
Phase space as a systems is described rather thoroughly, at any rate from the topological
symplectic manifold point of view. The points of a symplectic manifold (in mechanics it is
in mechanics called the phase space of a mechanical system) on which a completely
integrable system is given appear to split into three groups (types). Points
of type 1 are characterized by the fact that the integral trajectory sent from
them never returns to the same point but starts winding round a Liouville
torus which is its closure (Fig. 3.1.4). Points of type 2 are characterized
by the fact that the integral trajectory sent from them moves along the
Liouville torus and in some time goes back, i.e. the trajectory is closed.
In this case, the trajectory continues moving along itself winding on itself
infinitely many times. All the rest ofthe points are of type 3. We shall not
specify here the behaviour of the integral trajectory for points of type 3
since it may prove to be very complicated. We shall mention them later.
Non-resonant It turns out that for analytic non-resonant systems points of type 1 are
integrable systems typical, of general position and constitute the "majority" of all points.
Points of types 2 and 3 occupy a thin subset of measure zero.
Consider a point x of type 1 in a phase space M2n and denote by
Tn(x) the Liouville torus which is the closure of the integral trajectory
sent from the point x. When the point x changes, the torus m(x) also
undergoes some changes. Moving along the phase space, it can at some
instant degenerate into a circle. This happens, for example, if when mov-
ing the point x gets into the set of points of type 2. Thus, Liouville tori
may deform within the phase space transforming, splitting into a union
of other Liouville tori, merging with tori, degenerating, etc. What is hap-
Evolution and bifurcations pening with Liouville tori is important for understanding the geometry
of a Liouville tori of a given mechanical system since they show what is happening with
show the "topology the solutions of a Hamiltonian system of equations when the initial data
of a given system" are changed.
3.1 Some Concepts of Hamiltonian Geometry 199

In other words, the topological deformations and surgery on Liou-


ville tori demonstrate the qualitative character of the dependence of the
solutions of a Hamiltonian system on the initial data.
To the reader acquainted with mechanics we shall explain in addi-
tion that a point of a phase space is given by two sets of n number: n
coordinates and n momenta. Thus we have come to the natural problem
of classification of topological surgery on Liouville tori arising within Classification
a phase space. In the subsection to follow we shall specify the formu- of a transformations
lation of this problem. The mechanism forcing a Liouville torus drift of a Liouville tori
about a phase space should be described in more detail. Above we have
restricted our consideration to an arbitrary deformation of an initial point
x. But in physics and mechanics of particular interest are Liouville torus Evolution of
deformations caused by variations of the energy value. a Liouville tori
Let v be an integrable non-resonant Hamiltonian system on a sym- caused by variations
plectic manifold M2n. Let Hbe its Hamiltonian and 12, ... ,in additional of the energy value
independent smooth integrals in involution (including the Hamiltonian).
Consider an arbitrary Liouville torus. We state that it is specified only
by the Hamiltonian H = i1 and by the very fact of existence of other
additional integrals h,.'.. , in, i.e. a Liouville torus in general position Liouville torus in
does not depend on a concrete choice (concrete form) of the integrals general position does
h, ... ,in' To put it differently, if instead of the initially given integrals not depend on a choice of
h, ... , in one takes some other integrals 12,"" In (which as before are the additional integrals
independent of H and are in involution with it), then the Liouville tori
remain unaltered. This simple but important circumstance is explained
by the fact that a Liouville torus in general position coincides with clo-
sure of the integral trajectory sent from an arbitrary point and forming
an irrational winding on the torus. The integral trajectory itself, i.e. the
solution of the system is completely specified by the initial point and
by the Hamiltonian H and is absolutely insensitive to the choice of a
concrete form of the other integrals h, ... , in (they must only exist).
The very fact of existence of additional integrals implies that the tra-
jectory is winding round a torus (in non-resonant case). It should be
clarified here that additional integrals of the system are not, generally, Additional integrals are
uniquely determined. The functions 12, ... ,in can be replaced by some not uniquely determined
other ones, for instance, by those depending functionally on them and on
H. Instead of the set of integrals h, ... , in one can take, for example,
12 = 12 + 13'/3 = 12 - h h = Ii, 4 ~ i ~ n, etc.
200 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Momentum mapping 3.1.3 Momentum Mapping of an Integrable System

Let v be an integrable system with a set of independent involutive inte-


grals H = Ii, Jz, ... , In. Then with this system there is naturally associ-
ated an essential mapping, called the momentum mapping.
It is defmed as follows. Suppose F(x) = (fI(x), ... , In(x», i.e.
each point of the symplectic (phase) manifold is assigned a sequence
of numbers - the values of the integrals at this point. As a result,
there arises a smooth map of the manifold into a Euclidean space ~n
(Fig. 3.1.5). Setting a certain point a E ~n, w~ define the common level
surface Ta of all the integrals b, h ... , In. It is given by the equations
b(x) = ai, ... ,In(x) = an, where a = (ai, ... ,an), i.e. the numbers
ai are components of the vector (point) a in the space ~n. From the
implicit function theorem we know that the level surface is a manifold
Fig. 3.1.5 if all the integrals on it are independent. How can the independence of
the integrals be verified? To do this, one should first calculate their gra-
dients at points of the surface under consideration. Then the functions
b, h ... , In are independent in a neighbourhood of the point x if at
this point their gradients grad b, ... , grad In are linearly independent.
We recall that grad} = (a I / axi , ... , aI / ax2n ). At each point x of the
manifold M2n , the differential dFx of the map F is defined. Recall that
Critical points the point x is called critical for the map F if at this point the rank of the
of a momentum mapping differential dF is less than n, i.e. less than the dimension of the space
~n into which the manifold is mapped. The image F(x) of the critical
Critical values point will be called the critical value of the momentum map. The set of
of a momentum mapping all critical values will be denoted by E. It is sometimes called the bifur-
Bifurcation diagram cation set (the bifurcation diagram). Thus, if N is the set of all critical
points of the map F (in the manifold M 2n ), then E = F(N). The set E
is known to be the set of measure zero in the space ~n (the complement
of E is open and everywhere dense in ~n).

3.1.4 Surgery on Liouv1lle Tori at Critical Energy Values

Regular values Points a from ~n which are not critical values will be called regular
of a momentum mapping values (for a given momentum map). Let F-i(q) be a complete pre-
image of the point q under the map F. If a is a regular value and the
3.1 Some Concepts of Hamiltonian Geometry 201

pre-image p-l(a) is compact then, according to the Liouville theorem,


it consists of Liouville tori (Fig. 3.1.5). If a moves along a smooth curve
'Y in the space I.n and remains a regular value for the momentum map
then its pre-images deform smoothly in the manifold. In other words,
each of the Liouville tori Tn deforms smoothly inside the manifold M 2n Deformation of
and remains a Liouville torus. But the situation changes radically when a Liouville tori
a approaches the diagram E and at some instant of time punctures it. inside 2n-manifold
Let c be the point where the path 'Y meets the set E. Since c is the
critical value of the momentum map, its pre-image may be not a tours
(or a union of tori). To say it differently, it is precisely at this moment
that there occur surgeries (bifurcations) of the tori. Figure 3.1.5 illustrates Surgery of
schematically the surgery of tori at the moment when the path 'Y punctures a Liouville tori
the bifurcation set E.
We distinguish between two types of the sets E. Sets oftheftrst type Two types of
are characterised by the fact that their dimension is equal to n - 1. Sets of a bifurcation diagrams
the second type are such that their dimension is strictly less than n - 1. "large" and "small"
Note that since P is a smooth map, there always holds the inequality bifurcation diagrams
dimE:=;n-1.
Suppose for a given integrable system there holds the inequality
dim E < n - 1. In this case the pre-images p-l(a) and P-l(b) of any
two regular values a and b from I.n are "the same" in the sense that
they consist of one and the same number of Liouville tori, the manifold
p-l (a) being diffeomorphic to the manifold p-l (b). Indeed, to prove this
assertion it suffices to join the values a and bby a smooth path 'Y in I.n ,
which does not intersect E. This can be done since by the assumption
we have dim E :::; n - 2 (Fig. 3.1.6). Therefore, according to the implicit
function theorem, when the point 'Y(t) moves along the path 'Y from the
Fig.3.1.6
value a to b, its complete pre-image is deformed in the manifold M via
a diffeomorphism. As a result we obtain that each of the Liouville tori
sweeps a cylinder Dl x rn and that cylinders corresponding to different
tori do not intersect. This implies our assertion.
In this respect, the Hamiltonian systems in which dim E :=; n - 2 are
organized rather simply (form the topological point of view): any two
sets of Liouville tori of the form p-l(a) and P-l(b) can be deformed
smoothly into each other in the class of Liouville tori. Indeed, consider
corresponding points a and b and join them first by an arbitrary con-
tinuous path 'Y. It may generally speaking meet the bifurcation set E.
202 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

But since dim E ::; n - 2 and the path "{ is one-dimensional, one can
remove the path "{ from the set E using a small deformation. As a result,
we obtain the path joining the points a and b and not intersecting E.
Then the above arguments are valid. In this sense, when the "majority"
of Liouville torus deformations occur in the manifold M2n , the Liouville
tori undergo no surgery (and if some surgery does accidentally occur, it
can be removed by means of small "stirring").
We see quite a different picture when dim E = n - 1. Here the
set E, generally speaking, splits the space ]Rn into individual isolated
Chambers of "chambers" (Fig. 3.1.7). The point is that here E consists of pieces of
momentum mapping smooth hypersurfaces (perhaps with singularities). Therefore, E separates
]Rn. Consequently, choosing the points a and b in different chambers
we can see that any smooth path joining these points will necessarily
puncture the set E at some point c (or in several such points). Hence,
when moving along this path, a Liouville tours (in the manifold M) will
The case of "large" necessarily come across the set N of critical points and may undergo
bifurcation diagrams is a serious topological surgery. Moreover, any way we choose to send a
vel}' important in physics torus from F-l(a) into a torus from F- 1(b) will necessarily lead us at
some moment to a qualitative surgery.
Thus, if dim E = n - 1, then each chamber (i.e. each connected
-.' ----------- .--c::> component of the set ]Rn \ E) is generally characterized by its unique
~.~ A r i:: c:::::) set of Liouville tori. The number of these tori may be different for each
chamber. This situation is most frequently encountered in concrete inte-
-- ---- --- -- _.c::><:::::>
':. ::. 0 0
'"
grable systems in physics and mechanics. We shall therefore give some
o
more attention to the case dim E = n - 1.
o We shall discuss a concrete mechanism responsible for the drift of
"" real Liouville tori about a symplectic (phase) manifold. Consider a simple
example - a well with the profile shown in Fig. 3.1.8. Suppose that a ball
Fig.3.l.8 starts falling down along the well wall, the friction being now absent.
Then the ball motion is completely determined by its initial position (the
height), i.e. by its potential energy in the vertical gravitational field. In
the absence of friction the ball will each time reach the initial height, fall
Physical example down the well again, and so on. The process will be infmite.
Now we "switch on" the frictional force. Let the friction be very
small. Then we assume to a first approximation that the motion of the
3.1 Some Concepts of Hamiltonian Geometry 203

ball has not practically changed, and taking very small time intervals we
may say that the system behaves as before. But on large time intervals the Motion of a ball and
effect of small friction is telling appreciably. Clearly, in the course of time effect of small friction
the ball will rise at an increasingly small height due to the energy loss Decrease in the
(energy dissipation). Finally, the moment of qualitative transformation ball energy
of the ball motion will come: it will for ever remain either in the left or
in the right well (Fig. 3.1.8). The surgery proceeds at the moment when
the level of the ball rise touches the saddle A. If above the point A the
level line is homeomorphic to a circle, then at the moment the level line
touchs the saddle A (the critical energy level), it (the level line) becomes
a figure-of-eight, and as the energy goes on decreasing the level line
splits into two separete circles which go down until they finally become
the points of minimum [50].
Thus, the deformation of system's behaviour was caused here by
the decrease in the ball energy. In the general case this situation can be
described like this.
Consider an integrable system with integrals H = II, h, ... ,In. Fix General situation: the
the values of all the integrals but one, namely, let the values of the func- drift of Liouville tori
tions h, ... , In be fixed and the value of the energy integral (i.e. H) induced by change
vary. There starts the drift of the Liouville tori about the manifold, in- of energy value
duced by the variation of the H value. In the course of this, the torus
will sweep an (n + I)-dimensional surface which we denote by xn+l.
Clearly, it is the common level surface of the integrals h, ... , In (the
integral II is excluded). See Fig. 3.1.9. We may assume the surface X n+!
to be the pre-image of the straight line lying in ]Rn and passing through
the point a parallel to the coordinate axis H. The surface X may, ac-
cordingly, be considered outside the manifold M. Then H becomes a
smooth function on the surface, i.e. H can be regarded as the map of X
into a real straight line. As the energy increases, the Liouville torus or
tori corresponding to the energy value a start deforming on X. Then, at
a certain critical energy value c, the torus undergoes a surgery, i.e. splits
into several tori, and the latter, carried away by the increasing energy,
go on moving (Fig. 3.1.9).
The question is how the surgeries on Liouville tori are classified.
The answer for a rather large class of systems is given in the subse.ction
to follow. Fig.3.1.9
204 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

3.1.5 Visual Material

Hamiltonian systems occur in may mechanical and physical problems.


Motion of a rigid body Figure 3.1.10 illustrates the inertial motion of a rigid body in an ideal
in an ideal fluid incompressible liquid. The dynamics of such a system is described by
Hamiltonioan equations. This permits the use of the methods of algebra,
analysis and group theory in the study of such systems. In the case of
Hamiltonian field with 2-manifolds, a Hamiltonian field is depicted as a flow of incompressible
1 degree of freedom is liquid through a surface. In this case, incompressibility is equivalent to
incompressible flow area conservation by figures on the surface, carried along by a flow of
on 2-surface liquid (Figs. 3.1.1 I). The liquid may run off rather whimsically, but at
each nonsingular point of the flow on the manifold the divergence of
the flow is equal to zero. Figure 3.1.11 represents the front of moving
flow. Figure 3.1.12 presents mathematical variations of close themes of
optical effects, caustics, singularity theory [51]. These questions are also
connected with the theory of Lagrange manifolds [18].

Page 205, Fig. 3.1.1 o. Motion of rigid body in an ideal fluid

Page 206, Fig.3.1.11. Hamiltonian field on 2-surface as the flow of in-


compressible fluid. Atlantic myths: temple in the ocean

Page 207, Fig.3.1.12. Smooth 2-surfaces as a mirrors; optical and re-


flection effects. To M. Bulgakov's "Master and Margaret": Roman soldier
Mark controls the crucifixion of Jesus
208 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

3.2 QUalitative Questions of Geometric Integration of Some


Differential Equations. Classification of Typical Surgeries of
Liouville Tori of Integrable Systems with Bott Integrals

Nondegenerate 3.2.1 Nondegenerate (Bott) Integrals


(= Bott) integrals
In this section we present, in particular, the most visual aspects of the
new theory of Liouville torus surgery developed by the present author
[54], [55].
For simplicity we consider the case of a four-dimensional symplectic
Integrable Hamiltonian (phase) manifold with an integrable system on it. Although practically
systems with 2 degrees all the facts listed above remain valid in an arbitrary multidimensional
of freedom case too, we shall restrict our consideration to the case of dimension four
only to make it more demonstrative.
Handf- Here the integrable system possesses two independent integrals H
independent integrals and f, of which one is a Hamiltonian (the energy integral) and the other
is an independent additional integral which is in involution with H. The
Momentum mapping momentum mapping F maps the manifold M4 into a Euclidean plane
~2 for the coordinates of which we take the values of the integrals H
\~: :..,., :~--~~
-:."% __ -j__________
i and j, i.e. F(x) = (H(x), f(x». From the viewpoint of our problem the
::: C integrals Hand f are equivalent since they are in involution. Therefore
S -; ~ ~ H and f may exchange places according to which of them is of primary
X -<:7" --------- IX
interest in a given problem.
1 We shall now be concerned with a three-dimensional surface X 3 on
Fig. 3.2.1 which the integral f is constant. Restricting the Hamiltonian H to this
surface we obtain on it a new function which we denote by the same
letter not to introduce new notation. Since our task is to classify surgeries
on Liouville tori, we assume that as the energy H varies from the value
a to the value b(a < b) (the value of the second integral being fixed),
Transformation of a the Liouville torus undergoes surgery at some intermediate enery value
Liouville tori under the c (Fig. 3.2.1). This can be conveniently imagined as follows. Assume for
change of energy level simplicity that the part of the surface X 3 which is mapped by the function
H into the segment [a, b] to be a smooth three-dimensional manifold. To
do this, it suffices to require that the gradient of j be non-zero on this
part of the surface X. Then the function H is a smooth function on the
manifold X. The part of the surface X outside the segment [a, b] is of no
3.2 Qualitative Questions of Geometric Integration 209

interest for us now. The value c is the critical value of the function H on Singular level
the manifold x, i.e. on the level set Xc = {x EX: H(x) = c} = H-1(c) of energy function
there exist critical points of the functions H (at which gradH = 0).
Consider the case of general position, i.e. the case where the function H
has typical singularities. This circumstance can be specified. We shall
say that the smooth function H is a nondegenerate or Bott one on the Nondegenerate
manifold X if the critical points of the function H form nondegenerate or Bott integrals
critical submanifolds.
Let us decipher this definition. Consider the set of critical points of
the function H. It is required that this manifold be a smooth subman-
ifold in the manifold X, and not necessarily a connected one. More-
over, it is required that the Hessian d2H of the function H (the sec-
ond differential) be nondegenerate on planes orthogonal to the subman-
ifolds of critical points (Fig. 3.2.2). In other words, consider the matrix
J = (fiHjaxiax j ), where xi are local coordinates. Then it is required Fig. 3.2.2
that its kernel coincide with the tangent plane to the critical submanifold
and that on the orthogonal complement to this submanifold the matrix Bott integral
be non degenerate. This means that the restriction of the function H to is a Morse function
each normal disc of small radius with centre at a point on the criti- on a normal disc to
cal submanifold is a Morse function. Such functions were introduced a critical manifold
by M. Morse and then especially intensively studied in the well-known
papers by R. Bott. This is why we call them Bott functions. It was proved recently that
So, let the Hamiltonian H be a Bott function on the suiface X in any analytical integrable
the neighbourhood of the Liouville torus surgery of interest, i.e. in the system can be approximated
neighbourhood of the critical energy value. by an integrable system
We shall point out another visual interpretation of the condition of with a Bott integral
Bott character (the condition of general position). Consider the bifur-
cation set of a momentum mapping. From Sect. 3.1 it follows that of
particular interest is the case where E is one-dimensional, i.e. consists
of segments of smooth arcs on the plane. They may be glued together,
i.e. the set E has, generally speaking, some singularities. Surgeries of
Liouville tori occur at the moment when a point, moving along a curve
on the plane, punctures (at a point c) the bifurcation diagram (Fig. 3.2.3).
Clearly, we obtain a three-dimensional surface X3 (see above). Since Fig. 3.2.3
we are interested in the case of general position, we may assume that
the smooth segment 'Y punctures the curve E transversally at its non-
singular point (Fig. 3.2.3) and with a nonzero velocity. These conditions
210 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Almost an invesffgated single out the class of "typical segments" ,. Any segment can always
today physical integrable be transformed by means of small stirring into such a "typical segment".
systems are nondegenerate The exact formulation of these requirements is given above in terms of
(= with Bott integral) Bott (= nondegenerate) integrals.

3.2.2 Classification of Surgeries of Bott Position


on Liouville Tori

It turns out that typical surgery of Liouville tori can be effectively and
demonstratively described. There is a finite number of such basic surgi-
cal operations, and what is happening to a torus under surgery can be
represented in a three-dimensional space. Consider the following three
topological operations on a two-dimensional torus.
1) Consider a solid torus standardly embedded in 1R3 (i.e. a direct
product of a circle by a disc) (Fig. 3.2.4). Inside this torus consider the
axis, i.e. a circle. The boundary of the solid torus is a torus. Examine its
deformation. The torus contracts inside the solid torus and becomes the
axial circle. This surgery will be denoted by T2 ___ 8 1 --- O.
2.) 2) Drill from the solid torus a thin solid torus concentric to the
initial one, i.e. consider in 1R3 the standardly embedded thickened torus
- the direct product of the torus by a segment (Fig. 3.2.4). Consider the
"middle" torus in this thickened torus. We may construct a deformation
contracting the boundary of the thickened torus (Le. both the inner and
outer tori) onto the middle torus (Fig. 3.2.4). The two tori move towards
each other (one decreasing in size, the other increasing) and finally meet
in the middle of the thickened torus merging into one. This surgery will
be denoted by 2T2 ___ T2 --- O.
3) Drill two solid tori from the complete torus standardlyembedded
3
in 1R as shown in Fig. 3.2.4. Each of the inner solid tori goes once along
the axis of the large solid torus. We obtain a three-dimensional manifold
with boundary consisting of three tori. Now consider the deformation of
the outer torus. It contracts to two thin tori inside it (Fig. 3.2.4). At some
3)
critical moment, the outer torus becomes a surface whose orthogonal

~ "
t) b M
/.' : cross-section (by a plane orthogonal to the torus axis) is a figure-of-
eight. Then the figure-of-eight breaks down and splits into two circles.
This corresponds to a breakdown of the outer torus into two tori, after

29-0000
·. . . 6d.b.o Ag.3.2A
3.2 Qualitative Questions of Geometric Integration 211

which each of them contracts to its inner thin torus. We denote this
surgery by T2 ~ 2T2. It is quite clear that sending the arrows into the
opposite direction, we shall obtain from all the three indicated operations
their inverses.

Theorem 1. (Fomenko [54], [55]). Let v be an integrable (by means of


a Bott integralj Hamiltonian system on a four-dimensional symplectic
(phase) manifold. Then all the topological surgeries of Liouville tori in
Bott position that occur inside the manifold M4 (and are determined by
this system) split into compositions of three elementary surgeries of types
1,2, 3 described above.

The condition of Bott position implies that H is a Bott function


in the neighbourhood of its critical values (after the second integral on
the level surface is restricted). Theorem I admits a visual interpretation.
Consider a critical energy level sutface, i.e. a 2-surface H = c in a 3- Critical surface
manifold X (Fig. 3.2.5). Let us see what is happening to a Liouville torus for energy function
which starts moving (carried along by the variation of the energy value)
in the direction of this critical surface. Then in the general position only
three elementary events may happen to this torus (the other events are
compositions of the elementary ones).

I) The torus meets the critical surface H = c and having no strength


to break through it "damps" on it and becomes (degenerates into) a circle
(Fig. 3.2.5).
2) The torus moves towards the critical surface and having no
strength to break through it, strikes it and is reflected. After the stroke
and reflection, the torus again plunges into the region of lower energy
values.
3) The torus approaches the critical surface, strikes it, breaks the
Fig. 3.2.5
surface and splits into two tori which continue their independent upward
motion (already in the region of higher energy values).
Simple nondegenerate
Reversing the arrows (see also Fig. 3.2.5), we obtain the three inverse integrable system has
processes: exactly one critical
the torus falls from above onto the critical energy surface and fades manifold on each
to become a circle; critical level
212 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

the torus strikes the critical energy surface from above and is again
reflected in to the region of higher energy values;
two tori fall onto the critical surface from above, run into each other,
merge into one torus which breaking the surface continues its downward
motion.
We have described the elementary surgeries which form the "basis"
in the set of all surgeries of Bott position. Any typical surgery is a com-
position of elementary ones. It turns out that one can effectively describe
"second level" surgeries, i.e. "not very complicated" combinations of the
three above-mentioned surgeries.
Consider a critical energy surface and suppose there is exactly one
connected critical manifold of Bott Hamiltonian H on this surface. Then
while intersecting this critical surface a Liouville torus will undergo a
surgery. Let us call this surgery Hamiltonian and simple. Theorem 2
to follow provid~s a complete description of all possible Hamiltonian
(simple) surgeries. Before formulating the theorem, we point out two
more surgeries on the torus.
4) Consider a solid torus embedded standardly in 1.3 and going
twice along its axis drill a thin solid torus from it as shown in Fig. 3.2.6.
This gives a 3-manifold A3 whose boundary consists of two tori, outer
and inner. We shall describe the process of transformation of the outer
torus into the inner one. In Fig. 3.2.6 we see a deformation of the outer
Fig. 3.2.6 torus (in the normall cross-section of the solid torus, i.e. on a disc with
two holes). It contracts inside the solid torus and then transforms into
a torus wound twice along its own axis and tangent to itself along a
circle (Fig. 3.2.7). To obtain this surface, we should do the following.
We should take a standard flat circle, consider an orthogonal to it figure-
of-eight with loops labelled by numerals 1 and 2 and then start moving
the figure-of-height along the circle leaving the former orthogonal to the
latter but rotating it so that after a revolution it could occupy its initial
position but its loops 1 and 2 exchange places. The surface sweeped
by the figure-of-eight during its motion will just be the median position
of the outer torus which tends to transform into a thin one wound twice
round its axis. This surgery will be denoted by T2 - t T2. In our terminol-
ogy this surgery means the following. The torus approaches the critical
Fig. 3.2.7 energy surface (Fig. 3.2.7), breaks through it, and on transforming again
3.2 Qualitative Questions of Geometric Integration 213

into one torus goes on moving already in the region of higher energy.
Transformation of a torus into a torus consists in its double winding
round itself (the exact description is given above).
5) Consider a Klein bottle K2 immersed in ~3 (Fig.3.2.8). Al- Immersed Klein bottle
though this immersion has a circle of self-intersection, this is inessential as a critical level surface
for us here. Examine a tubular neighbourhood of this immersed subman- of a simple integral
ifold. This means that at each point x of the Klein bottle we take a small
segment centred at the point x and orthogonal to the surface. Uniting
all these segments, i.e. making the point x run through the whole of the
Klein bottle, we obtain a 3-manifold K3 immersed in ~3. This manifold
contains the Klein bottle and obviously contracts to it. For contraction
it suffices to make the points p and q (the end-points of the segment
centred at the point x, see Fig. 3.2.8) move towards each other in the
direction of the point x. As a result, the manifold K3 contracts to the
2-surface, the Klein bottle. Consider the boundary of the manifold K3. It
can be readily seen that it is diffeomorphic to a two-dimensional torus. It
is immersed in ~3 and in the course of the above-mentioned contraction
"covers" the Klein bottle. Given this, to one point there go exactly two
points of the torus. Namely, to the point x there go the points p and q.
To make sure that the boundary of K3 is a torus we recommend that
the reader draw carefully an immersed Klein bottle and trace out the
whole picture of contraction. An attempt to represent this picture was Fig. 3.2.8
made by the author in Fig. 3.1.19. In topology, such a transformation of
a torus into a Klein bottle is called a two-sheeted covering. We denote
this surgery of a torus by T2 - K2 - O. Here the Liouville torus ap-
proaches at low velocity the critical energy surface, does not break it but
"fades" becoming a Klein bottle that lies on the surface H = c.

Theorem 2. (Fomenko [54], [55]). Any Hamiltonian (simple) surgery ofa Simple integrable system
Liouville torus (of Bott position) coincides with one of the five elementary has exactly one critical
surgeries 1,2, 3, 4, 5 listed above. manifold on each
critical level
The Hamiltonian character of surgeries 1, 2, 3 is rather obvious. It
is more difficult to make sure that the Hamiltonian surgeries 2, 4 and 5
are compositions of the elementary surgeries 1 and 3. There holds the
following assertion:
214 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

a) A Hamiltonian surgery of type 4 is a composition of surgeries


Transformations 2, 4, 5 are 1 and 3; a surgery of type 2 is also a composition of surgeries 1 and 3;
topological compositions of b) A Hamiltonian surgery of type 5 is a composition of surgeries
the transformations 1 and 3 of types 1 and 4, i.e. the sum of two type 1 and one type 3 surgeries.

We shall not prove this fact here. For what follows it is convenient
to introduce visual notation. In Fig. 3.2.9 we show how we shall represent
surgeries 1-5 of Liouville tori. This will be also instructive in the study
of topological structure of constant energy surface.
So, surgery 1 (Fig. 3.2.9) means that the torus degenerates into its
axial circle. Surgery 2 (Fig. 3.2.9) shows a merging of two tori into one.
Surgery 3 (Fig. 3.2.9) is transformation of one torus into two. Surgery 4
(Fig. 3.2.9) is transformation of a torus into a torus twice wound round
the initial one. Surgery 5 (Fig. 3.2.9) implies transformation of a torus
into a Klein bottle (by means of a two-sheeted covering). Figure 3.2.9
illustrates the inverse surgeries as well. These are separated from the first
five by a dashed line.
If on one energy level there arise two or more surgeries, we shall
depict this event as shown in Fig. 3.2.9, where three types of surgery on
one energy level are given as an example.
3-t-3~ Clearly, from the elementary graphs of Fig. 3.2.9 we can compile
more complicated graphs by glueing the ends of the edges. The inte-
/\ Fig. 3.2.9
rior (nonsingular) point on the edge of the graph depicts schematically a
nonsingular Liouville torus. As the point slides along the edge the torus
moves inside the surface, on which the integral f is constant, sweeping
a part of it. At those points at which the edges of the graph merge or end
"Molecule" = as vertices, the tori undergo corresponding surgery. Thus, constructing a
one-dimensional graph shows one-dimensional graph, we draw the set of all Hamiltonian (simple) surg-
all topological bifurcations eries of Liouville tori (i.e. their decomposition into simple surgeries) on a
of Liouville tori 3-surface given by a constant value ofthe integral f. In Sects. 3.2.4-3.2.9
we employ this geometrical language for a visual interpretation of the
familiar results concerning topological surgery of Liouville tori in some
classical integrable physical systems.
Complete classification The complete classification of ALL surgeries of Bott position (i.e.,
of all complex (i.e. not not only Hamiltonian, simple) is given in the paper by A.T. Fomenko
only simple) bifurcations "Topological Invariants of Liouville-Integrable Hamiltonian Systems"
of Liouville tori published in Fllnktsional'nyi Analiz i eygo Prilozheniya, 1988, v.2,
3.2 Qualitative Questions of Geometric Integration 215

No.4, pp. 38-51. Here the integrable system can have many critical sub-
manifolds on the same critical level.

3.2.3 The Topological Structure of Critical Energy Levels at a


Fixed Second Integral

Consider a critical energy level H "" const ~n a manifold X 3, i.e. under Classification of all
the condition that the second integral f is fixed. The question is how this critical submanifolds
level is topologically organized. First we enumerate the possible critical of a nondegenerate
submanifolds of the energy integral on the surface X = (f = const). integrable systems

Theorem 3. (see [54], [55]). The critical points of a smooth Hamiltonian


H, which is a Bott function on a compact nonsingular manifold X3 =
(f "" const), occupy either isolated critical circles, or two-dimensional
tori, or Klein bottles.

Clearly, the Bott Hamiltonian H cannot have isolated critical points


on the compact nonsingular surface X3. We call the Hamiltonian H Simple and complicated
simple if on each of its critical levels there lies EXACTLY ONE con- (complex) Hamiltonians
nected critical submanifold. All the other Hamiltonians will be called
COMPLICATED. A small perturbation of the initial Hamiltonian Hand Any complex integrable
of the symplectic structure w on a 4-manifold proves always to be able system can be topologically
to transform any complicated Hamiltonian into a simple one. (This the- perturbated into a simple
orem has been proved by Nguen Tien Zung). Consequently, using an integrable system
arbitrarily small perturbation (of the Hamiltonian and of the symplectic
structure, or of the Hamiltonian and of the second, additional integral),
we can transform all complicated surgeries of Liouville tori into a set of
simple surgeries. At the same time, for a thorough qualitative analysis of
integrable systems it is exceedingly important to have the classification
not only of simple, but also of all complicated surgeries of Liouville
tori (i.e. without a perturbation of the Hamiltonian and integral). Such
a classification has also been obtained (see above).

Theorem 4. Let v be an integrable Hamiltonian system and let H be a


simple Bott Hamiltonian on a compact nonsingular sUlface X 3 "" (f =
const) (where f is the second integral of the system). Then each critical
216 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

energy-level sUrface on X. which is a smooth manifold, is necessarily


diffeomorphic either to a torus, or to a circle, or to a Klein bottle. Each
critical energy-level surface on X 3, which is not a manifold, is obtained
by glueing two two-dimensional tori TJ and T2 along non-contractible
cycles (circles) /'1 and /'2 located respectively in the tori TI and T2.
On such singular energy level, integral trajectories of the Hamiltonian
system wind asymptotically (in the case of Bott positon) onto this non-
contractible cycle or coincide with it.

Theorem 4 is illustrated in Fig. 3.2.10. We take, for instance, two


copies of a torus, on each of which a standard (non-contractible) parallel
is drawn. Glueing together the two tori along this parallel, we obtain a
surface whose cross-section is shown in Fig. 3.2.10.
Another example. We take two tori on each of which a standard
(non-contractible) meridian is drawn. Glueing together these tori along
the meridian, we obtain the surface shown in Fig. 3.2.10, where one of
Fig. 3.2.10
the tori is shown by a dashed line since in realization in 1.3 it is nec-
essarily inside the outer torus and is tangent to it from within along the
meridian. Figure 3.2.11 shows the behaviour of the integral trajectories
of an integrable system on a singular two-dimensional energy-level sur-
face (the value of the second integral is fixed). On each of the two tori
(Fig. 3.2.10) there exists a "locking" cycle on which the integral trajec-
tory is winding infinitely from the right and from the left. This cycle
Fig. 3.2.11 prevents the trajectories from revolving round the parallel (in the partic-
ular case of Fig. 3.2.11). This picture is seen to differ drastically from
Physical examples the case of rectilinear winding described by the Liouville theorem on a
of integrable systems nonsingular torus.

3.2.4 Examples from Mechanics. The Equations of Motion of a


Dynamics of a rigid Rigid Body. The Poisson Sphere. Geometrical Interpretation of
body in the 3-space Mechanical Systems

From now on, Sect. 3.2 demonstrates fruitfulness of the methods of vi-
sual geometry ~n the analysis of concrete mechanical systems. We employ
here the new language elaborated above and permitting, in particular, a
visual presentation, using the graphs r( Q), of some already familiar re-
3.2 Qualitative Questions of Geometric Integration 217

sults obtained by V.V. Kozlov, by Kharlamov [56], [57], Pogosyan and


Kharlamov [58], Tatarinov [59], A.A. Oshemkov and other authors. We
should mention here the important results by M.P. Kharlamov providing
a qualitative description of all integral level surfaces of many integrable
systems. In particular, the bifurcation diagrams proposed by M.P. Khar-
lomov proved to be exceedingly useful for a subsequent analysis of the
topology of integrable systems. Topology of
We assume a rigid body with a fIxed point to be in the fIeld of integrable system
axisymmetric potential forces. The configuration of a rigid body is the
motion of R3 transfonning a certain orthonormal frame into a fIxed or-
thonormal frame (el' e2, e3). Thus, the configuration space of a rigid Configuration space
body is the group SO(3), i.e. the group of proper rotations of a three- of a rigid body
dimensional space. Consider the one-parameter group 8 1 acting on the Action of 1-parametric
confIguration space M3 = SO(3) as follows: 8 1 = {"pT;T E R}, where subgroup on the space SO(3)
the transformations "p T are given by the formulas

"pT(Q) = ( 0I C?S0T - sin0)


T .Q , where Q E SO(3) .
o sm T cos T
This action defInes the rotation group around the vertical. We
assume the vector el to be directed vertically. We defIne the map

P(Q) ~ QT ( ~ ) ~ v E lII.', i.e. v ~ lbe vector composed of lb,


three vertical components of the frame, which under transformation with
the matrix Q becomes the frame (elo e2, e3). Obviously, vT +vi +v~ = 1,
and therefore the map p transforms the group SO(3) into the sphere
8 2 called the Poisson sphere. The pre-image of any point v E 8 2 Poisson sphere
(i.e. V[ + vi + ~ = 1) under the map p is exactly the orbit of the
action of the group 8 1. Hence, the Poisson sphere admits the repre-
sentation 8 2 = SO(3)/8 1• Since the potential II is symmetric about Potential II
the vertical axis, it follows that II = II (VI, 112, 11). The transforma-
tions tangent to the group transformations 8 1 can be shown to pre- Tangent bundle
serve the vector fIeld on T*SO(3) = SO(3) X R3, which detennines of the group SO(3)
the dynamics of the rigid body. Therefore, the dynamic system on
T*SO(3)/8 1= 8 2 X R3 = {vlv[ +vi + vj = I} x {W E R3}, where Wis
the angular velocity, is defIned. Angular velocity
218 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

For the axisymmetric potential II, this system is described by the


Euler-Poisson equation Euler-Poisson equation

{ (A,w) +w x Aw =v x gradII
v=vxw,

Inertia tensor v vi
where A is the diagonal matrix (inertia tensor), E 8 2 = {II[ + + vj- = I}
determines the position of the rigid body up to rotations about the ver-
tical, x is vector mUltiplication, (,) is the Euclidean scalar product of
vectors.
This system of equations always has two integrals, namely, the
Energy integral energy integral H(v, w) = 1/2(Aw, w) + II(v) and the area integral
Area integral G(v, w) = (Aw, v).
It can be shown that if there exists an integral K = K(v,w) indepen-
dent of the integrals G and H, then the connected component of the non-
singular surface Ih,g,k = {x E 8 2 x lR.3IH(x) = h, G(x) = g, K(x) = k}
is diffeomorphic to a torus T2, and the motion on it is conditionally
periodic.
The set of the critical values of the map I = H x G x K : 8 2 x lR.3 --+
Phase topology of lR. is the bifurcation diagram E C lR.3. To investigate the phase topology
3

a mechanical system of a mechanical system, it is necessary

1) to construct the bifurcation set E,


2) to indicate the number of tori T2 corresponding to each of the
components lR.3 \ E and
3) to examine the critical integral surfaces for the points of E.

Domain of possible motion The domain of possible motion (DPM) corresponding to (h, g, k) E
of a mechanical system lR.3 is the set Uh,g,k = P(h,g,k) C 8 2, where 8 2 is the Poisson sphere.
The generalized boundary of DPM of Uh,g,k is the set of critical values
of the map p = plh,g,k : h,g,k --+ Uh,g,k' A vector w E lR.3 such that
Admissible velocity (v, w) E h,g,k is called the admissible velocity at the point v E 8 2 for
the DPM of Uh,g,k'
The classification of DPM types with an indication of the set of
admissible velocities at each point is called geometrical analysis of a
mechanical system ([56]-[58]).
3.2 Qualitative Questions of Geometric Integration 219

3.2.5 An Example of an Investigation of a Mechanical System.


The Liouville System on the Plane Liouville system

The configuration space il,1 this problem is ]R2(XI, X2). The system is
defined by the Lagrange function: Lagrange function

( .2 .2)
'PI (XI) +'P2(X2)
L = 1/2 ()Ij (xd + >d X2» (XI (XI)xI + (X2(X2)x2 - Al (XI) +>d X2) ,

where Al (x]) + A2(X2)f 0, (XI(X]) > 0, (X2(X2) > 0. This system has two Integrals of
integrals: Liouville system

H = 1/2(AI + A2) ((XI xI + (X2X~) + ~: : ~: ' Lissajous figures

/ ( .2
K = 1 2(AI + A2) A2(X2 XI - AI(X2 X2 +
.2) 'PI A2-'P2 AI
Al + A2
Horizontal and
Consider the projection p : T*]R2 ---+ ]R2, p(XI, X2, XI, X2) = (XI, X2). Here vertical oscillations
instead of the Poisson bracket we take the configuration space ]R2. From
the conditions H = hand K = k it can be easily calculated that
.2 2('Pi - Aih +(-I)ik)
Xi = - (Xi(AI + A2)2 ' i = 1,2 .

Hence, Uh,k = P(h,k) = {(XI, X2) E ]R21Ai :S 0; i = 1,2}, where Ai(xi) =


Fig.3.2.12
'Pi - Aih + (_I)i k. Thus, the most typical version has the form U(h,k) =
[aI, bd x [a2, b2] C ]R2 (Fig. 3.2.12). For the interior points or'the rectan-
gle we have four admissible velocities: Two on its sides and one (zero) at
vertices. Trajectories in the rectangle (projections of phase trajectories)
are similar to Lissajous figures (the sum of independent horizontal and
verticale oscillations). The pre-image p-ICUh,k) is a torus with condi-
tionally periodic motion. It can be shown that the DPM type changes Fig. 3.2.13
in passing over (h,k), such that AI(xI) or A2(x2) has a multiple root.
Let xi be the multiple root of Al on a segment [al,bJ] (Fig.3.2.13).
Numerals in the figure indicate the number of admissible velocities. The
corresponding h,k is the direct product of a figure-of-eight by a circle
(Fig. 3.2.14). Its median line is the limiting cycle for the rest of the phase
(integral) trajectories. In a Liouville system, the DPM type changes only
for (h, k) E E. In the general case, the DPM type may change also
for those (h, k) which are not critical values of the momentum map
I=H x K. Fig. 3.2.14
220 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

3.2.6 The Liouville System on the Sphere

Let a rigid body be fixed in the centre of mass and let it be in the
central Newton field. Here JI(II) = (All, II) /2AIA2A3, and in addition
to the energy and area integrals there also exists the integral K(w, II) =
l/2(Aw,Aw) -l/2(A- I II,v). This integral was pointed out by Clebsch
for the problem of the motion of a rigid body in a liquid.
Consider the case G = O. We have:
TS2 = {(WII) E82 x 1R3IG(w, II) = O} .
Now iI = HT• S 2 and K = KIT.S2 are the integrals of the system on T*8 2•
Fig. 3.2.15 The bifurcation diagram of the momentum map I = iI x K : T*8 2 -+ 1R2
has the form shown in Fig. 3.2.15. The bifurcation set E consists ofthree
rays and a parabola segment tangent to two of them. The set E splits
the plane into five regions. For regions I, II, III we have h,k = 2T2 , for
region IV - h,k = 4T2, and for the other points, not lying on E, we
have h,k = 0. For the segments of E labelled by the numerall we have
h,k = 28 1, for segment 2 - h,k = 481 and for segments 3 - h,k = R X8 1
(Fig. 3.2.16). For segments 4 we have h,k = (8 1 V 8 1) X 8 1 (Fig. 3.2.14).
Next, the DPM for I has the form of two squares on opposite sides
of the Poisson sphere (Fig. 3.2.17). Projections of the phase trajectories
fill up the squares like Lissajous figures. Next, the DPM for type II
has the shape of a ring (Fig. 3.2.18), and for type III it is another ring
R.@ (Fig.3.2.19). On the ring, in both directions (since this is the image of
two tori) there go trajectories touching by tum the ring boundaries. Next,
the DPM for type IV consists of two rings (Fig. 3.2.20). Trajectories go
Fig. 3.2.16
in both directions on each ring. The surgery in going over from region
I to region II is shown in Fig. 3.2.21. Given this, the squares (DPM for
region I) are glued with their sides and form a "horizontal" ring (DPM
for region II). The surgery in going over from region II to region III
is shown in Fig. 3.2.21. Here the horizontal ring (DPM for II) is glued

Fig. 3.2.19

)It Pit-
Fig. 3.2.17 Vi

Fig. 3.2.18
Fig. 3.2.20
3.2 Qualitative Questions of Geometric Integration 221

along the segments of the cross-section V2 = 0, lying between the points


PI, P2 and P3, P4, and tears along the segments P2P3 and P4PI, where ~
PI, P2, P3, P4 are points on the cross-section V2 = 0 (Fig. 3.2.22). Here /\
Fig. 3.2.21
for the critical surface Lh,k the DPM is the entire Poisson sphere.
The surgery in going over from region III to region IV is represented
in Fig. 3.2.23. The ''vertical'' ring (DPM for III) breaks along the median
line. Transition from region I to region IV causes the surgery on Liouville
tori shown in Fig. 3.2.23. The corresponding events on the Poisson sphere
are depicted in Fig.3.2.24. So, the upper and lower sides of the square
are folded in two and glued, along the segment PI P4 there goes a break.
We obtain one of the rings DPM IV (Fig. 3.2.24). Fig. 3.2.22

- yy
Fig. 3.2.23

Fig. 3.2.24

3.2.7 Inertial Motion of a Gyrostat

Here the Euler-Poisson equations have the form


h c;][ ,.D
11 ~r
.AW +w X (Aw + A) = 0, iJ = v x w ,
i ,2. f<
where A is a constant vector in ~3. The integrals are of the form

H = 1/2(Aw, w), G = (Aw + A), v), K = (Aw + A), (Aw + A)) ; h


I =H x G x K : 82 x R3 ---+ R3 .

1) Let Al . A2 . A3 =t- O. The bifurcation set Eg = En {g = const}


is presented in Fig. 3.2.25. Here E splits the space R3 into four regions: Fig. 3.2.25
for region I we have h,g,k = T2, for II and III we have h,g,k = 2T2, and
222 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

for IV we have h,g,k = 0. The surgeries on the Liouville tori occurring


along the straight line A are depicted in Fig. 3.2.26. It is now easy to
describe to strucutre of Ih,g,k for points of the set E. For the reversal
points we have h,g,k = 8 1 x 8;, where 81 is a circle with angular point
Fig. 3.2.26 (Fig. 3.2.27).
2) Let )'1 = ).2 = ).3 = O. This is the Euler case for an ordinary

C>
rigid body. Given this, Eg degenerate as shown in Fig. 3.2.28. Except
for the trivial case 9 = h = k = 0, where h,g,k = 82 is a Poisson sphere
filled with motionless points, non-empty sets are obtained when k > O.
Fig. 3.2.27 Then one can introduce parameters c =l/k and (1 = 2h/k and project
the set E onto the plane (c, (1). This will give us the picture represented
in Fig. 3.2.29. Here the DPM type changes not only in passing over the
points (h, g, k) E E as in the case of Liouville tori in the preceding
examples. There are 25 DPM types altogether corresponding to the 25
regions into which the image on the plane (c, (1) is split. We obtain
;'=0
three basic types of images of the Liouville tori on the Poisson sphere
(Fig. 3.2.30, cases a, b, c). Cases a, b and c in Fig. 3.2.30 are distributed
It ii~_ .BL
on the plane (c, (1) as shown in Fig. 3.2.31. These images of tori are stable
," lI. in the sense of stability of smooth mappings. There may also exist two
",
'"
'!::"':, '1*0
JY
* types of unstable images (Fig. 3.2.30, cases d, e) realized on the curves
separating regions a, b, c.
Fig. 3.2.28

6" iv
tfAa .' .- .....
.. :'. Iff. . : '. Ii
l~ : ....... ."..:.
I" •• '.",." ." 8)
1 ......"" . "
~t "N'" ii e
Fig. 3.2.29 Fig. 3.2.31
Fig. 3.2.30

-1

Fig. 3.2.32
3.2 Qualitative Questions of Geometric Integration 223

3.2.8 The Case of ChapJygin-Sretensky

Here Al = A2 = 4A3, Al = A2 = 0, A3 = >., IJ(v) = -VI. In addition


to the energy integral H and the area integral G, there exists here a
wD
particular integral K = 2(W3 - A) (W[ + + 2W]/J3 on the set {G = O}.
Consider the map 1= fI x K : T.8 2 -+ 8 2. The bifurcation set 17(A) is
depicted in Fig. 3.2.32. The dashed lines do not belong to 17. Case "a" Fig. 3.2.33
is the Goryachev-Chaplygin case of the motion of a rigid body. The set
17(A) splits JR3 = JR2(h, k)XJRI(A) into seven regions. Consider the richest
case 4/3 < A2 < 4 (Fig. 3.2.32d). Here the plane JR2(h, k) is divided into (A)= l (C)1 (E)= III
seven regions which are the cross-sections of the corresponding seven (G)=tY (8)= f (~)= 111
f='JY
regions in the space JR\h, k, A) (Fig. 3.2.33). The surgeries on Liouville
tori along the arrows are given in Fig. 3.2.34. Here the surgery (H) is a (F)=lr l (H) =
composition of two surgeries of type (C) where both critical circles lie
on one level. The critical level surface is the direct product of a circle Fig. 3.2.34
by the curve obtained by glueing three copies of circles at two points.
We now consider the Goryachev-Chaplygin case (A = 0). The bifur-
cation diagram is presented in Fig. 3.2.35. Figure 3.2.36 shows a splitting
of the set {(h, k); k > O} c JR2(h, k) into ten regions, each corresponding
to its own DPM type. The dashed lines do not separate regions. DPM for
six cases J(a)- J(i) are depicted in Fig. 3.2.37 (see also Fig. 3.2.36). These
are images of one Liouville toms. One can see stereographic projections
from the pole (-1,0,0). Roman numerals indicate the number of admis- Fig. 3.2.35
sible velocities. Figure 3.2.38 shows DPM for four regions lI(a) - 1I(d).

Fig. 3.2.36

Fig. 3.2.38
224 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

This is the image of two Liouville tori. The projection of each of these
tori onto the Possion sphere is in this case symmetric about the plane
113 = O. Therefore, we depict here only the hemisphere 113 > O. The nar-
row band near the boundary of the ring in Fig. 3.2.38 is the image of one
of the tori· which lies entirely in the image of the other torus.

3.2.9 The Case of Kovalevskaya

Here Al = A2 = 2A3; >. = 0; II(v) = -VI. The additional integral


has the form K = (wf - w~ + VI)2 + (2WIW2 T v2i. Consider the map
I = H x G x K : 82 x ]R3 - t ]R3. The cross-sections of the bifurcation
diagram Eg are presented in Fig. 3.2.39.

a) g=O d) 1/2 < g2 < 4/3..;3


b) 0 < y2 < 1/2 e) g2 =4/3..;3
c) g2 = 1/2 f) 4/3..;3 < g2 < 1

.','
',' n.L _71"
......
:-:'

Fig. 3.2.39
3.2 Qualitative Questions of Geometric Integmtion 225

In region I we have h,g,k = T2, in II, III, IV we have h,g,k = 2T2,


in V we have h,k =4T2. For the surgery e.g. in the case 0 < g2 < 1/2
see Fig. 3.2.40.

8 (A)= 1(BJ=ll (C)=Y


(~)i?(E)=1 ! (F)=tt
Fig. 3.2.40

3.2.10 Visual Material

Figure 3.2.41 illustrates rotation of various tops differing from one an-
other by their moments of inertia, i.e. roughly speaking, by the form. The Moments of inertia
whirligig is one of such tops. A rapidly spinning top is one 6f the prin-
cipal parts of such wide-spread devices as gyroscopes. High demands Spinning top
are made of their symmetry and centring. If a top is non-symmetric, Gyroscopes
it can tear to pieces when it speeds up and the number of revolutions
is large. Gyroscopes are used for aligning various apparatus (airplanes,
ships, etc.). In modem mechanics of nonholonomic systems much atten- Nonho/onomic
tion is given to the study of solid bodies rolling along two-dimensional mechanical systems
surfaces as well as various "skates" or a sharpened plate (blade) sliding
on the ice or a similar surface. Integration of such equations of motion is
a fascinating but difficult task. See e.g. the recent papers by V.V. Kozlov
and Ya.V. Tatarinov. Figure 3.2.42 shows a Poisson spheres onto which
226 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

a three-dimensional manifolds and the comprised in it Liouville tori of


an integmble Hamiltonian system are mapped in the way described in
Sect. 3.2.4. The properties of the bifurcation diagram of momentum map-
ping can be investigated by tracing out the behaviour of images of the
Images of a Liouville torus on the Poisson sphere. In Fig.3.2.41 these images fill
Liouville tori on equatorial bands. Similar events are illustrated in Figs. 3.2.42. The images
the Poisson sphere of tori can be depicted as fairly exotic sets. Beside that, Figs. 3.2.43,44
Vortex motion illustrate vortex motions in liquid flows, surgery on level surfaces of
in the flow integrals of Hamiltonian systems. The inertial motion of a rigid body
in an ideal incompressible liquid is described by Hamiltonian equations.
But the appearance of turbulence immediately distorts and complicates
the patterns of the lines of flow. Various geometrical images associated
with such liquid flows (singular points of flows, vortices, sepamtrices of
flows, etc.) are presented here.

Page 227, Fig. 3.2.41. Gyroscops as a rapidly spinning tops; rolling ball
as a nonholonomic mechanical system

Page 228, Fig. 3.2.42. Poisson spheres and Euler-Poisson equation of a


motion of a rigid body. Ancient Egyptian priests and military court

Page 229, Fig. 3.2.43. Vortex motion in a flow

Page 230, Fig. 3.2.44. Boundary effects in the flow and analytic functions.
Greek and Atlantic legend about the daughter of the god Poseidon
3.3 Three-Dimensional Manifolds and Visual Geometry 231

3.3 Three-Dimensional Manifolds and Visual Geometry of


Isoenergy Surfaces of Integrable Systems

3.3.1 A One-Dimensional Graph as a Hamiltonian Diagram

We now consider constant-energy surfaces Q = Q3 (sometimes re-


ferred to as isoenergy surfaces), Le. surfaces described by the equa- Isoenergy 3-surfaces
tion H = const. We investigate nonsingular isoenergy surfaces, that is,
such surfaces on which the function H does not have critical points,
i.e. gradHrO everywhere on Q. The implicit function theorem implies
that in this case Q3 is a smooth 3-manifold in M4. Since M4 is always Isoenergy surfaces
orientable (as a symplectic manifold), so are the 3-surfaces Q. The term are orientable
"surface" should not be confused here with two-dimensional surfaces
which we studied above. It is a known fact of mechanics that equilib- Equilibrium positions
rium positions of a Hamiltonian system are given by the critical points of a Hamiltonian system
of the Hamiltonian (i.e. by such points where grad H = 0). Therefore,
when considering nonsingular (non-critic~l) constant-energy surfaces we
thus assume them to have no equilibrium positions of the system. We
also assume further on Q3 to be a compact closed manifold. Now restrict
the second integral f to the surface Q, and not to introduce new notation
denote the function obtained by the same letter f. Two-dimensional non-
singular level surfaces of the integral f on the manifold Q3 are obviously
Liouville tori. Varying the f value, we make these tori drift along the Liouville tori inside
manifold Q3. Clearly, the surgeries of Liouville tori (of Bott position) isoenergy 3-surface
caused by variations of the value of the integral f on the surface Q3
are to be described in an absolutely the same manner as those caused
by variations in the value of the Hamiltonian H on the surface X3 (see
Sect. 3.2).
We assume the integral f to be a Bott one on Q3. The Bott integrals Simple nondegenerate
can be divided into two Classes: simple and complicated. We shall the (= Bott) integrals
integral f simple if each connected component of the critical level sur-
face rl(c) contains exactly one connected critical submanifold of the
integral. The other Bott integrals will be called complicated. In other Complicated (= complex)
words, the integral is complicated if on some connected component of nondegenerate integrals
the level surface f-1(c), where c is the critical value of the integral f,
there lie several critical submanifolds.
232 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Simple nondegenerate We shall now, be concerned with simple Bott integrals. For the case
integrals of complicated integrals see the papers [60], [61] and the book [62] by
the author.
Almost all known It turns out that the overwhelming majority of known physical and
physical integrable systems mechanical integrable systems possess Bott integrals, and are non-
are nondegenerate resonance on almost all nonsingular Liouville tori. A Hamiltonian system
Non-resonant systems is called non-resonance (on an isoenergy. surface Q) if the Liouville tori
on which the trajectories of the system form an irrational winding are ev-
erywhere dense in a given isoenergy manifold Q3. It turns out (for details
see Ref. [62]) that if a Hamiltonian system is non-resonance (on some
Q), its topology does not depend on the choice of the second integral f
and is completely determined by the Hamiltonian H itself.
We find it convenient to determine to topology of a manifold Q (and
at the same time the behaviour on it of any second integral f independent
Graph as a diagram of the Hamiltonian H) by means of a graph which we denote here by
of an integrable system r(Q, H, f) or simply by r(Q, H) since, as mentioned above, in the non-
resonance case this graph does not depend on the choice of the integral
f. We describe the construction of this graph in the case of a simple
integral f (more precisely, in the case of those Hamiltonian H which
admit at least one simple integral f). The construction of a graph in
the case of a Hamiltonian admitting complicated integrals is given in
Ref. [62].
A graph r(Q, H) will be called the diagram ofan integrable Hamil-
tonian H on a given isoenergy surface Q = (H = const).
The integral f can be viewed as a smooth map of the manifold Q
into a real straight line lR 1• If a is a regular value in lR 1, then the pre-
image of the point a under the map f consists of a finite number of
two-dimensional tori. Depict each torus as a point. Varying the value a
we make these points move. As a result, they sweep the edges of some
___________ fR.1
graph. If the value a has become critical, the edges can merge or, on
~ _ --
_ Gt the contrary, split into several edges, or run into a vertex. We therefore
obtain a certain graph r(Q, H, f) which we call the diagram of the
simple integral f on Q (Fig. 3.3.1). Clearly, this graph shows not only
the topological structure of the manifold Q, but also the main details
of the behaviour of the integral on Q since the merging and splitting
Fig. 3.3.1 of edges and other singular points (vertices) of the graph correspond to
distinct critical manifolds of the function f on Q.
3.3 Three-Dimensional Manifolds and Visual Geometry 233

The theorems formulated in Sect. 3.2 allow us to say a good deal


about the geometry of the graph r. The point is that we have in fact Structure of the graphr
described the structure of all critical submanifolds of the simple Bott in the case of a simple
integral f and, therefore, we already know all the forms of the splitting nondegenerate system
and merging of the edges of the graph and all the types of its vertices.

Theorem 1. (see Refs. [54], [55]). Let Q3 be a compact non-singular


isoenergy surface of an integrable system with a simple Bott integral
f. Then all possible types of vertices of the graph r(Q, H, f) and all Five types of "atoms" =
possible types of interaction of its edges are listed in Fig. 3.3.2. In par- vertices of graph r
ticular, the vertices of the graph are divided into five types. All the
above-mentioned types of vertices and interactions of edges are actually
realized in real mechanical and physical integrable systems.

Vertices of types 1, 2, 5 depict critical manifolds of the integral f


on Q, which are minima or maxima of the integral. Such vertices and
the corresponding submanifolds are called minimax. Vertices of types 3 Minimax and saddle
and 4 depict the saddle critical submanifolds of the integral. We shall atoms represent different
accordingly call them saddle vertices. Furthermore, one can say exactly types of transformations
which particular critical manifolds are depicted by these or those vertices of Liouville tori
of the graph.

1) The black-minimax vertices of the graph are minimax circles of


the integral.
2) The white minimax vertices of the graph are minimax two- "Atoms" are represented
dimensional tori. as a 3-manifolds with
3) The tripods (see type 3) are critical saddle circles of the integral boundary tori
with the property that when passing through them a Liouville torus splits
into two tori (or, inversely, two tori merge into one as the direction of Atom A =minimax
motion reverses). atom B = saddle = tripod
4) The star with one incoming and one outgoing edges of the atom A' = "star"
graph is the critical saddle circle of the integral in passing through which
a Liouville torus again transforms into one torus. This transformation
consists in two-fold winding round the torus.
5) The white minimax vertices of the graph with a point inside
them are minimax Klein bottles.
234 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Now we are in a position to demonstrate more clearly the processes


~'"
i 1
where a torus "strikes" the critical energy level (these processes are

.:. :.::,/;:.::. described in Sect. 3.2). See Fig. 3.2.2. We can see that in cases I, 2, 5

7t7 ..:.::*
........
'::. ':;"
the Liouville torus does not break into the critical level of the integral
f for the simple reason that this level is either maximal or minimal.
"Above" the maximum and "below" the minimum there are no points
of the manifold Q at all, and therefore the torus has nowhere to move.
As a result, in case I the. torus becomes a minimax circle and stops. In
case 5 the torus transforms into a Klein bottle and stops. In case 2 two

Fig. 3.3.2

tori merge becoming a minimax torus and stop, or else we may assume
the torus to rise from the depth of the manifold, to strike the minimax
level of the integral and, reflected, to go again inside the manifold. It
is only in cases 4 and 5 that the critical levels of the integral are inside
the manifold Q. When passing through the type 3 level, the torus splits
into two (or tori merge into one) and continues moving in the same
direction. In case 4 the torus, on transforming, also continues moving in
r
Graph is called molecule the same direction. In Fig. 3.3.3 we have schematically shown the graph
(constructed from atoms) r(Q, H, f) of the general form (for a simple integral) where all possible
types of vertices are represented. From the viewpoint of the problems in
/R.1
question we may think that all the minima of the function f are at one
level (e.g. f = 0). Similarly, all the maxima of the function f may be
thought of as lying at one level, for instance, f = I (Fig. 3.3.3).

3.3.2 What Familiar Manifolds are Encountered Among Isoenergy


Surfaces?

In a three-dimensional space we consider a closed orientable 2-surface


M;, i.e. a sphere with ghandles. Suppose a heavy material point is
Fig. 3.3.3 moving along this sphere. There arises a mechanical system whose con-
3.3 Three-Dimensional Manifolds and Visual Geometry 235

figuration space (i.e. position space) is the surface M2. The phase space Phase space
is a 4-manifold T*M 2• Its points are paris (x, ~), where x is a point
of the surface and ~is a tangent vector to the surface at the point x
(Fig. 3.3.4). An important example of a Hamiltonian system on the space
T*M is a geodesic flow of some metric. To determine this system, a Geodesic flow
Riemannian metric 9ij should be introduced on the surface M. Then we
can calculate the distance between any two points on the surface. On
the surface there arise geodesic lines, i.e. lines minimizing the distance
between any two of its sufficiently close points [63]. For example, such
shortest (i.e. geodesic) lines on the Euclidean plane are straight lines.
On a sphere embedded standardly in ]R3 (i.e. endowed with a standard Fig. 3.3.4
invariant Riemannnian metric) the geodesics are the various equators,
i.e. flat cross-sections of the sphere by planes passing throu~h the centre
(Fig. 3.3.5). A geOdesic flow is a vector field on the space T*M defmed
thus. Consider an arbitrary point (x,O on T*M. It defines uniquely the
point x on M and the tangent vector ~ to M at this point. It is known
that from any point of a manifold one can always send a single geodesic
,et) in an arbitrary direction. Consider a point ,et) sliding along the
geodesic I (as a function of varying time t) with the velocity vector
'Y(t). We obtain the curve aCt) = <'Y(t), 'Y(t)) in the space T*M. In doing
so, we assume x = ,et), ~ = 'Y(t). So, we obtain a family of curves
Fig. 3.3.5
aCt) = (x(t), W)) = (,(t), 'Y(t)) on T*M. Considering their velocity vec-
tors we obtain a smooth vector field on T M. This is just the geodesic Geodesic lines
flow of the given metric. Changing the metric, we, generally speaking,
change the geodesic flow. Let us consider the matrix G = (gij), where
G-l = @ij), i.e. G-l is the inverse matrix. Tangent bundle T. M
It turns out that on the space T* M a symplectic structure can be is diffeomorphic to the
introduced in a natural way, and the manifold T* M can thus be trans- cotangent bundle T* M
formed into a symplectic one. Given this, the geodesic flow appears to Geodesic flow always is a
be a Hamiltonian field. It always has the energy integral H written as Hamiltonian vector field
H(x, 0 = 9ij(X)~i~j, i.e. it coincides with the scalar square of the tan-
gent vector ~ at the point x measured relative to the Riemannian metric
9ij on the manifold. From the mechanical point of view, the function
H(x,O is interpreted as the kinetic energy of the point moving on the
surface. In a frictionless motion of a material point along a surface in the
absence of external forces, its kinetic energy is conserved. In the motion
along geodesics, the length of the velocity vector of the geodesic remains
236 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Hamiltonian H(x, 0 unchanged and, therefore, the function H(x, {) is conserved along the in-
and kinetic energy tegral trajectories of the geodesic flow on the space T*M. This means
that H(x, {) is the integral.
Thus, any geodesic flow always has a first energy integral H(x, ~).
Geodesic flow of Consider the simplest example of a geodesic flow. If as the manifold

j
Euclidean straight line M we take a one-dimensional Euclidean straight line ]Rl with the metric
ds 2 = dx 2, where x is a Euclidean coordinate, then the phase space T*M
/R1 J is naturally identified with the plane ]R2 defined by Cartesian coordhlates
J tp.1 x and ~. Then the geodesic flow of the Euclidean metric on the straight
line will be depicted as a vector field on the plane ]R2, described by
e)
x -f-f-f-f++-H~ the equation v = (x, = (1,1). See Fig. 3.3.6. The integral trajectories
o of this flow are straight lines parallel to the bissecting line of the first
coordinate angle.
If as the manifold M we take a circle in a Euclidean plane, the

e--····~· phase space T*SI is identified with a straight circular cylinder.·1t is


..- ......... coordinatized by the variable cp (which changes along the circle, i.e. is
'" an angle) and the variable ~ (along the vertical generator ofthe cylinder).
'. See Fig, 3.3.6. On the circle we consider a locally Euclidean metric ds 2 =
dcp2. Then'the geodesic flow of this metric will be depicted by the vector
Fig. 3.3.6 field on the cylinder. The integral trajectories of the flow are helical lines
winding round the cylinder (Fig. 3.3.6).
Geodesic flow of Consider the geodesic flow of a Riemannian metric gij on a surface
a Riemannian metric M;. As we know, it always has the energy integral H(x,~). Therefore,
on a 2-surface the isoenergy surface of this Hamiltonian system may be regarded as
the set of points (x, {) satisfying the condition H(x,~) = a2 = consti' O.
What is the geometric representation of the constant-energy surface?
Recall that H(x, {) determines the square of the length of the vecor
Isoenergy 3-surface
~ tangent to the surface at the point x. Consequently, if the point x
of a geodesic flow
is fixed, the equation H = a2 describes (providing the coordinates are
on a 2-surface
appropriately chosen) the circle in the tangent plane TxM to the surface
Mat the point x (Fig. 3.3.7). We denote this circle by S!. Its equation
will be 9ij(x)~i~j = a2. Since, generally speaking, the functions gij(X)
vary from point to point, the radius of the circle S! will be distinct at
different points. Here we have used the fact that the kinetic energy of
a material point is written as a positive definite nondegenerate bilinear
form on tangent vectors to the surface. This is particular implies that the
radius of the circle S! is always nonzero at all points of the surface.
3.3 Three-Dimensional Manifolds and Visual Geometry 237

Thus, the isoenergy 3-surface Q in this example is a compact 3-


manifold given by the equation H(x,~) = a2 in the space T*M. What
can be said about the topology of this manifold? The principal property
of this surface is that there exists a natural map (projection) 7r of the
manifold Q3 onto the manifold M2. Namely, 7r : Q --t M, 7r(x,~) = x.
Clearly, this projection covers the entire surface M. Which points from
Q are mapped into a single point x from M? It is clear that these points
form the circle S! (Fig. 3.3.7). In this case the map 7r is said to determine
a fibre bundle with base M and a circle as fibre. The manifold Q fibres Fig. 3.3.7
over the surface M with a circle as fibre. So, we can formulate the
following theorem.

Theorem 2. The isoenergy surfaces Q3 of the geodesic flow of a Rie- Isoenergy 3-surface
mannian metric on a two-dimensional compact closed surface M2 are is the fibration: Q3 ~ M2
three-dimensional compact manifolds fibred over M2 with a circle as
fibre.

Which of the above-mentioned geoQesic flows are integrable and


which are not? In our case the geodesic flow is depicted as a field Integrability and
on the phase 4-manifold T*M. It always has one integral, namely, the nonintegrability
Hamiltonian H(x, ~). Therefore, for a complete integrability it suffices to of geodesic flows
have one more integral f independent of H (consequently - in involution on 2-surface
with H). If such an additional integral does exist, the integral trajectories
of the geodesic flow "lie" on two-dimensional Liouville tori. A great
many papers are now being devoted to the problem of the existence of
a second integral, but we shall not list them here.
It is intuitively clear that the answer may depend on the proper-
ties of the Riemannian metric and on the topology of the manifold M. Some obstacles
The situation with integrability and nonintegrability in geometry and me- to integrability are
chanics has now been substantially clarified. To illustrate this, we offer produced by the topology
the following table. For details see e.g. Refs. [27], [21], [28], [50] as well of a manifold
as some other contributions by J.D. Birkhof, A. Poincare, V.V. Kozlov,
S.L. Ziglin, V.N. Kolokoltsov, Va. V. Tatarinov, S. V. Bolotin, I.A. Taima-
nov and others.
The Table implies that an obstacle to integrability of an analytic
geodesic flow may in any case be the topology of the manifold: in
238 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Analytical geodesic flow case there are many handles, the integral does not exist (y.V. Kozlov
of a 2-sphere with many [21], [50]). In case there are few handles (the sphere, the torus), the main
handles is non-integrable role is played by the analytic properties of the metric. In some cases the
integral does exist, while in others it does not. The metrics for which the
second integral does exist can be described qualitatively (although this
is not a simple task).
Weare now concerned with the question of what the isoenergy
surfaces in integrable systems are. On answering the question we shall
Topological obstacles understand the topological obstacles for integrability. For example, ifwe
for integrability manage to describe to class of isoenergy surfaces of integrable systems
more or less efficiently, then on discovering that in some concrete me-
chanical system a given isoenergy surface does not belong to this class,
we can immediately declare that the system is non integrable (at least on
the given isoenergy surface).
As we can see from the Table, for obtaining concrete examples we
should consider simple metrics on simple manifolds, i.e. on the sphere
and on the torus. We begin with a sphere endowed with a standard metric
invariant under rotations. The Hamiltonian H is given here as H(x,O =

Simple manifolds: Complicated mani-

~
sphere and torus folds: spheres with
more 'than two
metnc handles
Question: is it true that, Simple Riemannian Most often the If the metric is ana-
if geodesic flow of a metrics, e.g. pos- system is inte- lytic, the system is
smooth metric on a 2-torus sessing a large grable always nonintegrable,
is integrable, then this symmetry group is i.e. there no se-
cond independent in-
flow always has a quadratic
volutive integral
(in momenta) second integral? Complicated In some cases the If the metric is ana-
Riemannian integral exists, lytic, the system is
metrics in some cases not. always noninte-
There are criteria grable
(though not always
effective) for fin-
ding out if there
exists a second
integral
3.3 Three-Dimensional Manifolds and Visual Geometry 239

(the Euclidean length of the vector ~ tangent to the sphere at the point The next question: is it true
x). The geodesic flow of such a standard metric on the sphere is known that any "integrable smooth
to be completely integrable, i.e. there exists a second integral. The reader metric" on a 2-torus
may try to construct it. The isoenergy surfaces Q3 have here the form is equivalent to a
r
Q = {(x, 0, where I~ I= a = const O}. Therefore, Q is formed by points Liouvilfe metric?
of the form (x, ~), where ~ is a tangent vector of length a at the point
x E 8 2 (Fig. 3.3.8).

Lemma 1. Isoenergy surfaces Q3 of an integrable geodesic flow of a


standard metric on the sphere are diffeomorphic to the three-dimensional
projective space lRp3.

Proof. We associate each pair (x, 0, where I~I = 1 (we may assume
without loss of generality that a = 1), with an orthogonal frame in lR3 in
the following way. We assume that the sphere has a unit radius, trans-
port the tangent vector ~ parallel to itself to the centre of the sphere and
complement the pair of vector x and ~ with a third vector y so as to
obtain the fram e(x,~) with the same orientation which was fixed be-
forehand in lR 3 (Fig. 3.3.8). With each pair (x, ~) we associate the frame
e(x, O. Inversely, each orthogonal frame uniquely defines a pair of the
Fig. 3.3.8
form (x, ~). Thus, we have proved that the manifold Q is homeomor-
phic to the set of all orthogonal frames (with origin at the point 0)
in a three-dimensional space. It remains to prove that each orthogonal
frame is uniquely defined by a point of the projective space lRp 3. It is a
well-known fact of linear algebra that the set of orthogonal frames of a
fixed orientation in R3 is identical to the set of orthogonal proper trans-
formations of lR 3, i.e. to the orthogonal matrix group SO(3). We shall
prove that the group SO(3) is homeomorphic to lRp3. Each orthogonal
Remark: isoenergy 3-surface
transformation A: lR 3 -- lR 3 is uniquely determined by some motionless
of any geodesic flow on
axis l around which we observe rotation through an angle rp (Fig. 3.3.8).
a 2-sphere (not necessary
Plotting on the straight lihe l the angle of rotation, we obtain the point
integrable) is diffeomorphic
(l,rp) in lR3. If rp is smaller than 7r, the point uniquely determines the to lRp3
initial orthogonal transformation. If the angle of rotation is equal to ± 7r,
we obtain one and the same orthogonal transformation of lR3 although
the two points (l, +7r) and (l, -7r) representing it are distinct in lR3. These
two points should be glued, i.e. thought of as one point. So, the set of all
orthogonal (proper) matrices is homeomorphic to the following object.
240 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Class (H) of all "integrable" One should take in 1.3 a three-dimensional ball restricted to a sphere of
isoenergy 3-surfaces radius 1[' and glue diametrically opposite points on its boundary (i.e. on
has several different the sphere). This is the well-known model of a projective space. This
equivalent representations implies the lemma.
Class (H) coincides with We denote by (H) the class of all isoenergy surfaces of integrable
the class (W) of all (by means if a Bott integral) Hamiltonian systems. The set (class) of all
graph-manifolds introduced compact closed orientable 3-manifolds will be denotes by (M). It is clear
by Waldhausen that (H) C (M).
Lemma 1 implies that the projective space Rp3 belongs to the class
(H). Consider a torus T2 and endow it with a locally Euclidean (flat)
metric. This can be done since the torus is represented as the quotient
space of the Euclidean plane by the action of the group Z EEl Z. It is a
known fact that the geodesic flow of a flat metric on a torus is completely
integrable. Repeating here the procedure used in the preceding case, we
see that the isoenergy surface Q3 is homeomorphic to the set of pairs of
e
the form (x, e), where x is a point on the torus and is a tangent vector
of length 1 (at this point).

Lemma 2. The isoenergy surfaces Q3 of the integrable geodesic flow


of the standard flat metric on the torus are diffeomorphic to the three-
dimensional torus T3.

Proof. Consider a torus as a square with its sides identified (Fig. 3.3.9).
At each point of the square consider the tangent circle (of radius I) on
the plane. Clearly, their orientations on the glued sides of the square are
compatible. This implies that Q is homeomorphic to the direct product
of the torus by a circle, as required, since T2 = 8 1 X 81.

Hence, the three-dimensional torus also belongs to the class (H).


In mechanics there are examples of well-known integrable systems
whose Hamiltonian H reaches an isolated minimum or maximum at some
point of a symplectic (phase) space M4. Consider such a minimax point
Fig. 3.3.9 of a Hamiltonian, i.e. the equilibrium position of a system. Let H be
for example equal to zero at the point Xo of the minimum. Consider the
constant-energy surface Q given by the equation H = c > 0, where c
is sufficiently small. Since H(xo) = 0 and the point Xo is an isolated
3.3 Three-Dimensional Manifolds and Visual Geometry 241

minimum, the equation H(x) = c describes a three-dimensional sphere Remark: isoenergy 3-surface
8 3 centred at the point Xo. Thus, the standard three-dimensional sphere of any geodesic flow on
8 3 belongs to the class (H). a 2-torus (not necessary
Finally, mechanics gives us examples of integrable Hamiltonian sys- integrable) is diffeomorphic
tems on n1anifolds M 4, whose Hamiltonian reaches either maximum or to a 3-torus
minimum on the circle 8 1• Such are, for example, some systems de-
scribing the dynamics of a three-dimensional heavy rigid body (fixed at
one point). As in the preceding example, we again consider the level
surface of a Hamiltonian H, close to the minimax circle. Clearly, it is
homeomorphic to the direct product of the circle by the sphere 8 2• The
point is that at each point x of the critical circle, the disc normal to the
circle is a 3-manifold. Consequently, the equation H = const determines
the 2-sphere in this 3-manifold (= disc). Transporting the point along the 3-sphere as an "integrable"
circle, we obtain the required direct product. Thus, the manifold 8 1 x 8 2 isoenergy 3-surface
also belongs to the class (H). Let us sum up the results.

Proposition 1. To the class (H) oJisoenergy surfaces oJintegrable non-


degenerate (BotO Hamiltonian systems there belong, in particular, the 51 X 52 as an "integrable"
Jollowing manifolds: 83, lRp3, T3, 8 1 X 82. isoenergy 3-surface

What unites these manifolds from the topological point of view? It is


now difficult to answer this question because it is unclear in what terms
it should be formulated. It turns out (and this will now be explained) Theorem: class (H) does not
that all these manifolds are represented in the form of glueing simple contain hyperbolic 3-manifolds,
elementary "bricks". So, there arise some elementary manifolds which thus all hyperbolic manifolds
may subsequently be glued in an arbitrary order to form new 3-manifolds. lie outside of (H)
All the 3-manifolds obtained by glueing these bricks appear to belong to
the class (H) and, moreover, the class (H) contains no other manifolds
(this is a nontrivial theorem, see Refs. [54], [55], [64]). We proceed to the
study of the elementary bricks employed for construction of isoenergy Class (H) coincides with a
surfaces of integrable systems. class of all 3-manifolds
obtained by glueing of a
Proposition 2. A1l3-manifolds enumerated in Proposition 1 are obtained solid tori and some number
by glueing two solid tori (through some diffeomorphisms oj their bound- of copies of 51 x N2, where
aries which are tori). N2 is a disc with 2 holes
242 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Proof. Consider two solid tori PI, P2 and let TI and T2 be their bound-
aries homeomorphic to a torus. Let I: TI ---t T2 be an arbitrary torus-
to-torus diffeomorphism. On each of these tori we determine standard
parallels and meridians denoting them respectively by 0:1, (31 and 0:2, (32
(Fig. 3.3.10). The diffeomorphism maps the curves 0:1 and (31 into some
smooth curves /(0:1) and /((31) on the torus T2. In the preceding chapter
we got acquainted with deformation of closed curves on a torus. Recall
that by deforming continuously the curves /(0:1) and /((31) we can make
them become compositions of the basis curves 0:2 and (32. For example,
the curve /(0:1) is always homotopic to a curve of the form 0:~(3~, i.e. one
should first go b times round the meridian f3z and then a times round the
parallel 0:2. Similarly, the curve 1((31) can be deformed into a curve of
Fig.3.3.10 the form o:~(3r There arises a square integer-valued 1* = (~ :). If
Group of unimodular we consider only those diffeomorphisms which preserve torus orientation
integer-valued matrices (i.e. those fixed beforehand), this matrix is unimodular, i.e. thas a unit
is isomorphic to the group determinant, ad - be = 1. So, each diffeomorphism 1 of tori is assigned a
of al/ homotopic classes of matrix 1* ..Inversely, with respect to each unimodular integer-valued ma-
diffeomorphisms for the torus trix A one can construct such a diffeomorphism 1oftori that 1* coincides
with the matrix A. Glueing two solid tori via the diffeomorphism 1 of
their boundary tori, we obtain a closed compact 3-manifold. Since a dif-
feomorphism is determined by a matrix, it follows that each unimodular
integer-valued matrix A is assigned a certain 3-manifold M = M(A). As

an example we consider the matrices ( _ ~ ~) and (~ ~) .


3-manifolds: 53, lR'p3, r 3, Glueing them by means of a solid torus, one can readily make sure that
51 x 52 are obtained by this gives respectively the sphere 8 3 and the manifold 8 1 x 8 2• We leave
glueing of two solid tori it to the reader as an exercise to construct matrices generating lRp3 and
T3.

Thus, the above-mentioned isoenergy surfaces are obtained by glue-


ing two copies of one elementary brick - a solid torus (via a diffeomor-
phism of the boundaries).
3.3 Three-Dimensional Manifolds and Visual Geometry 243

3.3.3 The Simplest Isoenergy Surfaces (with Boundary)

Consider the following five three-dimensional manifolds with boundary.


Type I A solid torus, i.e. the direct product 8 1 x D2. Its boundary
is one torns (Fig. 3.3.11).
Typell The direct product T2 x DI will be called a cylinder. Its
boundary consists of two tori (Fig. 3.3.11). A cylinder can be realized as
a region in 1.3 (Fig. 3.3.11). To this end one should drill from a solid
torns a thin solid torns concentric to the initial one.
Type III The direct product A3 = N 2 X 8 1 will be called "trousers"
or an oriented saddle, where N 2 is a disc with two holes (Fig. 3.3.11).
To realize N. 2 x 8 1 in 1.3, one should drill from a solid torns two thin
solid tori. The boundary N 2 x 8 1 consists of three tori.
Type IV. Consider a solid torns in 1.3 and drill from it a thin solid
torns going twice round the axis of the initial one (Fig. 3.3.11). This
gives a 3-manifold whose boundary consists of two tori. We call it a
non-oriented saddle (non-oriented trousers).
Type V. Take a Klein bottle immersed in 1.3 (Fig. 3.3.11) and con-
sider its tubular neighbourhood K3. As we have already found out, the
boundary of this neighbourhood isa torns immersed in 1.3. The tubu-
lar neighbourhood is formed by small segments orthogonal to the Klein
bottle and having the centre at its points. This gives the 3-manifold K3
immersed in 1.3. To remove its self-intersections, it suffices to pass over
from 1.3 to 1.4 . Owing to the appearance of another dimension, one can
remove by slight stirring all self-intersections of the Klein bottle (recall
that the set of its points of self-intersection forms a circle) and, therefore,
remove self-intersections of the 3-manifold K3. So, K3 can be smoothly
embedded in 1.4 . We call K3 a non-orientable cylinder.

Lemma 3. (see Refs. [54], [55]). Only two of those five manifolds are
topologically independent. Manifolds IL IV and V are glueings of man-
ifolds I and III Namely, it can be formally written that IV = I + I I I
andY = I + IV, i.e. V = 2· I + III. This means that A3 is obtained by
glueing a solid torus and trousers, while K3 is obtained by glueing two
solid tori and trousers. Finally, I I = I + I I I. Fig.3.3.11
244 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

The ''plus'' sign indicates here the glueing of elementary bricks via
some diffeomorphisms of their boundary tori. We shall prove, for exam-
ple, the ftrst relation. Consider manifold IV realized in 1.3• Consider the
centre of a 2-disc (with two holes) and a circle h which passes through
this centre and is the axis of the large solid torus (form which we have
drilled a thin solid torus going twice along the axis). From the large
solid torus we discard a small tubular neighbourhood of the circle h, i.e.
drill a thin solid torus concentric to the large solid torus. We ftbre the
remaining part R into circles (the dashed lines in Fig. 3.3.12) which go
twice along the axis of the large solid torus. On the disc we draw two
radii and label them by the letter a with arrows (Fig. 3.3.12). It is readily
seen that R becomes the direct product of the disc with two holes by the
circle, i.e. is trousers.
Similar arguments should be applied to manifold V. We omit the
proof and leave it to the reader.
It can be shown that all the five manifolds enumerated above are
realized as pieces of some isoenergy surfaces of classical mechanical
Fig. 3.3.12 integrable systems. Therefore we have called them isoenergy surfaces
with boundary.

Topological classification 3.3.4 Any Isoenergy Surface of an Integrable Nondegenerate


of isoenergy 3-surfaces of System Falls into the Sum of Five (or Two) Types of Elementary
integrable non-degenerate Bricks
systems
Theorem 3. (Fomenko [54], [55]). Let M4 be a smooth symplectic
(phase, compact or noncompacV manifold and let v be a Hamiltonian
Liouville foliation system with Hamiltonian H, Liouville-integrable on one non-singular
of a symplectic manifold compact three-dimensional isoenergy surface Q, by means of a Bott in-
M4 does not depend tegral f. Then the critical submanifolds of the integral f on Q can be
on the choice of a only as follows: 1) minimax circles, 2) minimax two-dimensional tori, 3)
second integral f saddle circles with orientable separatrix diagram, 4) saddle circles with
non-orientable separatrix diagram, 5) minimal two-dimensional Klein
bottles. Let m,p, q, s, r be respectively the numbers of these critical man-
ifolds. Then the manifold is represented in the form of glueing (via some
difJeomorphisms of boundary tori) of the following elementary bricks:
3.3 Three-Dimensional Manifolds and Visual Geometry 245

Q = mI +pII +qIII +sIV +rV; the elementary 3-manifolds I, II, III,


IV and V are described above.

This theorem provides a visual topological classification of isoenergy


surfaces of integrable Bott systems. The numbers m,p, q, s, r have a clear
interpretation - they point out how many critical manifolds of each type
the integral f has on Q.
If we forget for some time about a concrete form of the integral
and conern ourselves with the simplest topological representation of the
isoenery surface, the answer will be given by the following theorem.

Theorem 4. (see [54], [55]). Let Q be a compact nonsingular isoenergy Theorem: any compact
surface of a Hamiltonian system integrable by means of a Bott integral orientable smooth 3-manifold
f. Then Q can be represented in the form Q = m' I +q' II I, where m' and is an isoenergy surface for
q' are some non-negative integers related to the numbers from Theorem some Hamiltonian system
2 by the formulas: m' = m + s + 2r + p, q' = q + s + r + p. (not necessary integrable)

So, each isoenergy surface of an integrable system admits two de- Hamiltonian decomposition
compositions: Hamiltonian (into the sum of five types of bricks) and and topological decomposition
topological (into the sum of two types of bricks). The first decompo- of isoenergy surface
sition is a more "detailed" one, it "remembers" the structure and the
number of critical manifolds of the integral f. The topological decompo-
sition is more rough, though it is simpler (as far as notation is concerned).
It ignores minute properties of the integral. Clearly, the Hamiltonian de-
composition is uniquely restored from th~ diagram r(Q, H, f) of the Classes (H) and (a) coincide
simple integral f.
Consider the various closed compact 3-manifolds obtained by glue- Connected some of two
ings of elementary bricks of types I and III. Denote this class of 3- "integrable" isoenergy
manifolds by (Q). The theorems formulated above show that (H) C (Q). 3-surfaces is again an
The inverse inclusion is valid too. It turns out (Brailov and Fomenko isoenergy 3-surface for
[64]) that the classes (H) and (Q) coincide, i.e. any 3-manifold obtained some integrable system
by glueing solid tori and trousers can be realized as a compact isoenergy
surface of an integrable (by means of a Bott integral) Hamiltonian system
on an appropriate symplectic manifold M4 (may be noncompact).
246 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

3.3.5 New Topological Properties of the Isoenergy Surfaces Class

Recall that a manifold M3 is called topologically reducible if it is rep-


resentable as a connected sum M = M, UM2, where M, and M2 are
not nomeomorphic to the standard sphere 8 3. Otherwise the manifold is
called irreducible. The connected sum is defined as follows. From each
of the manifolds M, and M2 we discard a ball. This gives manifolds with
2-sphere as boundary. Glueing these two spheres via an identity diffeo-
morphism, we obtain a new manifold denoted by M, UM2 (Fig. 3.3.13).

Theorem 5. (A.T. Fomenko, H. Zjeschang). Let Q, and Q2 be isoenergy


Fig.3.3.13
surfaces from the class (H) = (Q). Then their connected sum Q, UQ2
If v and v' are two integrable again belongs to this class. Inversely, if the isoenergy surface Qfrom
systems on isoenergy surfaces the class (H) = (Q) is represented as the connected sum Q, UQ2 of some
Q and QI, then we can always manifolds Q, and Q2, then each of them belongs necessarily to the same
construct a new integrable class (H) = (Q).
system v Uv' on the
connected sum Q UQI It is not only the manifolds themselves but also the integrals on them
that decompose here into a connected sum. Thus, the class of isoenergy
surfaces of integrable Bott systems is closed with respect to the opera-
tion of taking a connected s~. Now we are ready to answer the ques-
tion whether there exist purely topological obstacles for integrability of
a Hamiltonian system on an individual isoenergy surface. Since we are
dealing with smooth systems it is not excluded that the system may be in-
tegrable on one isoenergy surface and nonintegrableon another. It would
New topological obstacles to be desirable to have invariants forbidding (or allowing) the existence of
integrability of a Hamiltonian a Bott integral only on a single isolated isoenergy surface irrespective of
dynamical system the behaviour of the system on other isoenergy surfaces (including the
closest ones). It turns out that such topological obstacles do exist. We
shall not go into detail here and restrict ourselves to visual geometrical
interpretation. For details we refer the reader to Refs. [54], [55].
Class (H) of all "integrable"
isoenergy 3-surfaces is less Theorem 6. (see Refs. [54], [55]). The class of isoenergy surfaces of
than class (M) of all compact integrable (by means of a Bott integralj systems does not coincide with
orientable 3-manifolds the class of all three-dimensional compact closed orientable manifolds.
3.3 Three-Dimensional Manifolds and Visual Geometry 247

Corollary. Infinite subclass of3-manifolds cannot be isoenergy surfaces If a Hamiltonian system v has
ofBott integrable systems, i.e. there exist topological obstacles forbidding an isoenergy surface which does
integrability of a Hamiltonian system if a given isoenergy surface does not belong to the class (H).
not belong to the class (H) = (Q). then v is nonintegrable;
for example: any Hamiltonian
To solve efficiently the question of whether the Hamiltonian system system is non-integrable on
is integrable on a given isoenergy surface, one should be able to calculate hyperbolic closed 3-manifold
efficiently the topological obstacles briefly described above. To this end
it is useful to know as much as possible about the topology of the class of
integrable isoenergy surfaces. We shall describe some of their interesting
properties.
Suppose a smooth 2-surface M is homeomorphic to a sphere and
endowed with a smooth Riemannian metric ofgeneral position, i.e. on this
surface there is not a single stable closed geodesic. Then the geodesic
flow corresponding to this metric is non integrable on each individual
nonsingular isoenergy 3-surface in the class of Bott integrals [54], [55].
We shall discuss the concept of stability of a periodic solution of
a system v on an isoenergy 3-surface. The periodic solution is depicted
as a closed integral trajectory 1 (Fig. 3.3.14). We consider a tubular Fig. 3.3.14
neighbourhood of the trajectory, i.e. a solid torus whose boundary is
a 2-torus with the axis I' We say that the trajectory 1 is stable if the
solid torus fibres wholely (without "gaps") into concentric 2-tori with the
circle 1 as their common axis, all these tori being invariant with respect
to the system v. This means that all integral trajectories of the system
which are close to 1 lie on the indicated 2-tori.
The property of the system to have stable periodic solutions appears For any analytic integrable
to be closely connected with its integrability. There holds the following system (M4, H, f) always
assertion: it on an isoenergy surface the system has not more than one there exists a small smooth
stable periodic solution and moreover if the one-dimensional integer- perturbation such that a new
valued homology group HI (Q, Z) is a finite cyclic group (i.e. has the integrable system ( M4 , A, f)
form Zq), then the system is non integrable on this isoenergy surface is Bott system on a given Q3
(it cannot have a smooth Bott integral). Roughly speaking, if on the (V. V. Kalashnikov, jn.)
isoenergy surface there are many nonzero cycles offinite order and few
stable periodic solutions, the system is necessarily nonintegrable. See
also [65].
So, the very fact of the existence of a Bott integral f in the system
imposes strong restrictions upon the topology of the isoenergy surfaces. It
248 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

Set of Bott systems on a is natural to ask whether these restrictions remain if the requirements on
fixed isoenergy 3-surface the integral are weakened. For instance, we require from the function f
is of the first category in the "less" than in the preceding cases. Consider the class (8) of 3-manifolds
set of all smooth integrable with the property that on them there exists a smooth function 9 in which
systems; consequently form all the critical points are organized into nondegenerate critical circles and
a thin set in weak topology all the nonsingular level surfaces are unions of 2-tori. In other words,
the manifold belongs to (8) if and only if on this manifold there exists
a Bott function in which all the critical manifolds are circles and all
the nonsingular level surfaces are tori. Given this, the function does not
already have necessarily to be an integral of some integrable system.
Bottsystems are dense in Nevertheless, the class (8) appears to coincide with the class (Q) (S.V.
the set of all integrable Matveev) [66]. Thus, if on the 3-manifold M there exists a Bott function
smooth systems which have whose critical manifolds are circles and the non-critical level surfaces
critical circles only and consist of tori, then on this manifold there necessarily exists an integral
exactly one such circle on f of some integrable Hamiltonian system and the manifold is realized as
each critical level for H its isoenergy surface. Thus, (H) = (Q) = (8).
In some situations the requirements on the function f can be weak-
ened still more. Consider the class (R) of 3-manifolds on which there
exists a smooth Bott function h all of whose critical manifolds are cir-
cles. As distinct from manifolds of the class (8), it is not required here
that the nonsingular level surfaces of the function be tori. In the literature
such functions are called round Morse functions (Thurston, Azimov and
others) [67]. See also the paper by Matveev, Fomenko and Sharko [68].
Four faces of the remarkable Clearly, (R) :J (H), i.e. any isoenergy surface of an integrable sys-
class (H) of "integrable" tem admits a round Morse function. It can be proved that the class (R) is
isoenergy 3-surfaces: strictly larger than the class (H), i.e. there exist manifolds which admit
(HJo=(QJo=(SJo=(RJo a round Morse function but are not realized as isoenergy surfaces.
and (H)=(Q)=(S) The distinction between the classes (R) and (H) is due to their
different behaviour under the operation of taking the connected sum of
If "integrable" isoenergy manifolds. According to Ref. [67], if M is any 3-manifold (we do not
surface is a connected sum of specify each time that we work in the class of compact connected closed
some manifolds K and N, then orientable manifolds), then while considering its connected sum with a
both K and N are "integrable" sufficient number of copies of the manifold 8 1 X 8 2 we finally necessarily
get to the class (R). In other words, if m is sufficiently large, we have
M ~ (~f,!1 8 1 X8 2) E (R) . As has already been mentioned, the class
(H) possesses the remarkable property that if M = MI ~ M2 E (H) then
the manifolds MI and M2 belong necessarily to (H). In terms of the
3.3 Three-Dimensional Manifolds and Visual Geometry 249

class (8) this property has beeen proved by S. V. Matveev. From this
it immediately follows that the class (R) is strictly larger than the class
(H). Indeed, it suffices to take the manifold Mo which does not belong to
the class (H) (such manifolds are known to exist). Taking its connected
sum with a sufficiently large number of copies of the manifold 8 1 x 8 2,
we get to the class (R). The manifold M obtained does not however lie
in (H) since otherwise the initial manifold Mo would belong to the class
(H), which contradicts the choice of Mo. Thus, ME (R), but does not
belong to (H).
The picture changes radically if we restrict our consideration to the
class of irreducible 3-manifolds. It turns out that inside this class there
holds the remarkable identity: (H)o = (Q)o = (8)0 = (R)o, where "0"
indicates the set of irreducible manifolds.
This means that given an irreducible 3-manifold with a round Morse
function h on it, it is stated that on the same manifold there necessar-
ily exists also a Bott integral of an appropriate integrable Hamiltonian The orbital (continuous)
system, and the manifold will be an hmenergy surface of this system. classification of integrable
The topological theory of integrable systems, which we have briefly systems with two degrees
described here, was further developed by the author, in particular, in of freedom was obtained by
Refs. [60], [61], [62]. For example, the author discovered the topologi- Bolsinov and Fomenko in 1993
cal invariant classifying the integrable systems up to rough topological (see Matem. Sbornik, in print)
equivalence. The fme classification (up to fine topological equivalence)
was completed by the author in collaboration with H. Zieschang, S.V.
Matveev and A.V. Bolsinov [69], [70].
It turns out that many interesting aspects of three-dimensional topo- Bolsinov and Fomenko in 1993
logy [71], [72] are closely connected with Hamiltonian mechanics [73]. obtained the exact (continuous)
classification of a smooth
Hamiltonian systems on
2-surfaces (1 degree
3.3.6 One Example of it Computer Use in Symplectic Topology of freedom)

We have got acquainted with the remarkable class (H) of isoenergy 3-


manifolds of integrable Bott Hamiltonian systems. According to Chap. 2
(see Sect. 2.5), 3-manifolds can be ordered in the increasing order of their
complexity, and there exists only a finite number of irreducible manifolds
of a given complexity. This fact permits computer use for studying the Computers in 3-topology
topology of "sufficiently simple" 3-manifolds. Developing further the and in symplectic geometry
250 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

investigations carried out by the author [54], [55], S.V. Matveev obtained
the following result.

Any closed orientable Theorem. All closed 3-manifolds of complexity not exceeding 8 belong
3-manifold of complexity :::; 8 to the class (H), i. e. are isoenergy surfaces of integrable nondegenerate
is an "integrable" (Bott) Hamiltonian systems.
isoenergy 3-surface
This fact has been proved using computer and is based on the results
of Chap. 2 and on the classification theorem for isoenergy surfaces of
integrable systems, proved in Refs. [54]; [55].
At first Matveev proved this theorem for the manifolds of complexity
~ 5. Let us demonstrate the rough idea of this proof.
We shall analyze the belonging of all closed manifolds of complexity
~ 5 and almost all manifolds of complexity 6 to the class (H).

I) A 3-manifold M will be called toric if its boundary consists of


Toric 3-manifolds several tori or is empty. Let M be a toric 3-manifold of complexity k, i.e.
let M have an almost special spine with k vertices. Using simplification
technique (see Ref.[34]) one can prove that there exist toric manifolds
Mi such that:
1. Mi has a special spine with ~ k vertices;
2. M is obtained from Mi by means of the followings oper-
ations: glueing the trousers N 2 X 8 1, taking sums connected with one
another and with D2 x 8 1, 82 X 8 1 or RP3. Making these operations,
Complexity of a 3-manifold is we do not overstep the limits of the class (H). Therefore, if all toric
the number of vertices in its manifolds with special spines with ~ k vertices lie in (H), so do all toric
minimal almost special spine manifolds of complexity ~ k.
II) All toric maI\ifolds having special spines with zero or one vertex
lie in (H). There exist three such manifolds. From Ref. [46] it follows
that all toric manifolds of complexity ~ 1 lie in (H). Note· that there
exists a toric manifold of complexity 2 not lying in (H) - this is the
complement of figure-of-eight.
Filtration of 3-manifolds III) Let P be a special spine of a closed 3-manifold M and let 0: be
with respect of complexity one of its 2-components. Puncture it, i.e. discard an open 2-disc from it.
After a collapse we obtain an almost special spine PI of a toric manifold
MI, from which M is obtained by glueing a solid torus. The number of
vertices kl of the spine PI is less than the number of vertices k of the
3.3 Three-Dimensional Manifolds and Visual Geometry 251

spine P since in the cource of collapse the vertices on the boundary of


the component a and, perhaps, some other vertices vanish. For example,
if a adjoins twice a certain edge, there vanishes another 2-component
adjoining the same edge. We call the 2-component a large if k ::; l.
IV) Looking through all the special spines with::; 5 vertices, yielded Thurston'S conjecture:
by computer, shows that all of them have large 2-components. Not all the volume of hyperbolic
of the special spines with 6 vertices have been counted. But those that 3-manifold 01 is minimal in
were, also had large 2-components. The presence of a large 2-component the class of all hyperbolic
guarantees the belonging of the manifold to the class (H), since it is ob- closed 3-manifolds;
tained by glueing a solid torus to a manifold from (H). Let us summarize V91 01 ~ 0.98 ...
the results of this computer experiment. but it turned out that the
volume of O:! is less!
There exists a natural filtration of 3-manifolds by complexity, i.e. vol O:! ~ 0.94 ...
by the number of vertices in the minimal almost special spine [46].
For each k there exists only a fmite number of irreducible 3-manifold
of complexity ::; k. The analysis of the results of enumeration of 3-
manifolds of complexity ::; 5 carried out, using computer, by Matveev
& Savvateev [46] and then by Matveev for complexity ::; 10 shows that
all closed orientable 3-manifolds of complexity ::; 8 lie in the class (H).
Closed orientable 3-manifolds which do not belong to the class (H)
appearfor the first time in complexity = 9. We shall denote them by Q,
and Q2. The first of them appeared to be known to Thurston [74], [75],
and the second was discovered independently by J. Weeks (76] and by
S.V. Matveev & A.T. Fomenko [48].
Both these remarkable manifolds proved to be hyperbolic (i.e. man- New conjecture: O:! has
ifolds of constant negative curvature), the second (the one discovered by the smallest volume in class
Weeks, Matveev and Fomenko) having the smallest volume of all closed of all closed compact
hyperbolic 3-manifolds known today. Matveev and Fomenko hypothe- hyperbolic 3-manifolds
sized (and grounded their hypothesis) that this manifold Q2 is actually
a manifold of the smallest hyperbolic volume in the class of all closed Two remarkable hyperbolic
hyperbolic 3-manifolds. 3-manifolds 01 and O:!
For the details of this investigation see S.V. Matveev and A.T.
Forrienko "Constant Energy Surfaces of Hamiltonian Systems, Enu-
meration of Three-Dimensional Manifolds in the Increasing Order of
their Complexity and Calculation of the Volumes of Closed Hyperbolic
Manifolds" published in Russian Math. Surveys 1988, Vol.43, issue 1,
pp.3-24.
252 3. Visual Symplectic Topology and Visual Hamiltonian Mechanics

New theorem: geodesic flows Another remarkable property of these two manifolds is that if a
of two ellipsoids in]R3 are Hamiltonian system has an isoenergy 3-surface homeomorphic either to
orbitally (continuous) Q1 or to Q2, this system is not integrable on this 3-surface (in the class
equivalent if and only if their of Bott integrals). This reveals new topological obstructions for integra-
semiaxes are proportional bility, that lie in the topology of the manifolds Q. These obstructions are
(Bolsinov, Fomenko) essentially different from all familiar topological obstructions for inte-
grability of Hamiltonian systems.
It turns out that the following striking result is true (see [48]):
If the 3-manifold Q3 is hyperbolic, then any Hamiltonian system is
non-integrable (in the class of Bott integrals) on Q3. In other word, the
class (H) of isoenergy 3-surfaces of integrable Hamiltonian systems does
not contain hyperbolic manifolds.

3.3.7 Visual Material

One ofthe elements in Fig. 3.3.15 presents an oriented separatrix diagram


of the critical circle of a Bott integral. It is formed by segments of the
integral trajectories stuck to the circle. Two versions are possible: the
diagram is homeomorphic either to a ring cut along its axis by a critical
circle (Fig. 3.3.15) or to a Mobius strip. These two rings (strips) move
apart and along (up and down) the manifold. Figure 3.3.16 is basically
aimed at providing psychological perception of geometrical infinity.

Page 253, Fig. 3.3.15. Separatrix diagram of the critical manifolds for dy-
namical systems. German and Scandinavian legends; death of Alberich
and punishment of the gods

Page 254, Fig.3.3.16. Mathematical infinity and singular points. Medita-


tion and dreams of ancient priests
4. Visual Images in Some Other Fields of Geometry
and in Its Applications

4.1 Visual Geometry of Soap Films. Minimal Surfaces Soap films

4.1.1 Boundaries Between Physical Media. Minimal Surfaces Minimal surfaces

Suppose in 1.3 there are several adjoining but not mixing physical media,
for instance, in a large vessel there are several immiscible fluids. Suppose,
the whole system is in equilibrium. Since the media are immiscible, the
boundaries (interfaces) between them are determined. These interfaces Interfaces between
can be tought of (in the first approximation) as two-dimensional piece- phYSical media
wise smooth surfaces separating the adjoining media. We consider for
simplicity the case of two media which we denote by Al and A2• Let the
pressures in the media be respectively equal to PI and P2. The equilib-
rium condition for the media proves to impose a strong restriction upon
the geometry of their interface. To formulate this restriction, we require
an important concept of local differential geometry, namely, the concept
of mean curvature of the surface.
Let P be a point on a smooth two-dimensional surface M2 in 1.3.
At the point P consider a unit normal n, i.e. the vector of unit length
orthogonal to the tangent plane to the surface at the point P (Fig. 4.1.1).
Draw a plane II through this normal. It crosses the surface along some
smooth curve 'Y. We are here interested only in a small neighbourhood
of the point P. So, we obtain a flat curve lying in the plane II. It is a
known fact that at each point of a curve its curvature can be calculated.
Geometrically, the curvature is expressed as the inverse quantity to the
radius of the so-called adjoining circle, i.e. the circle tangent to the curve Fig.4.1.1
256 4. Visual Images in Some Other Fields of Geometry

at the given point and approximating the curve to within small quantities
Function of curvature of second order (second-order tangency). The curvature can also be in-
terpreted from the mechanical point of view. Along a curve line we send
a material point of unit mass with a velocity constant in the absolute
Principal curvatures value. The direction of the veloctiy vector will, of course, change (it is
>'1 and >'2 are equal to directed along the tangent of the curve), but we assume its length to be
the maximal and minimal constant. We thus obtain a mechanical system whose state is determined
values for the curvature by the shape of the curve. When moving along the curve, the material
of a normal cross-section point has the change the direction of motion, due to which it accelerates.
This acceleration is depicted as the vector orthogonal to the curve (more
precisely, to the tangent to the curve at the given point). The magnitude
of the acceleration is just the curvature of the curve (at a given point).
The curvature (acceleration) may change from point to point. We obtain
a smooth function called the function of curvature of a flat curve. The
curvature of a spatial curve is defined in a similar way.
The curvature of a straight line is identically zero: a material point
moves along a straight line uniformly and rectilinearly without accel-
eration. The curvature of a circle of radius R is equal to 1/ R. The
acceleration is here constant in the absolute value.
Let us return to the surface M. To each position of the plane II
(passing through the normal to the surface) there corresponds its own
intersection curve·, and, therefore, its own value of the curvature k(,)
of this curve at the point P. Rotating the plane II (around the normal),
Mean curvature = >'1 + >'2, we change the value k(,). Consider the maximal value Al of this cur-
Gauss curvature =>'1 . >'2 vature and its minimal value A2. We obtain two numbers. They may, in
particular, coincide. From the definition we see that the curvature k{[)
is constant in this case, i.e. independent of the angle of rotation of the
plane.
Consider the general case. The numbers Al and A2 are called the prin-
cipal curvatures of the surface at a given point. These numbers change
from point to point, i.e. Al and A2 are functions of the points.
The sum of the principal curvatures, i.e. the expression Al + A2 is
called the mean curvature H(P) of the surface at a given point.

Laplace-Poisson theorem The Laplace-Poisson theorem. Let M be a two-dimensional boundary


between two media in ]R3, which are in equilibrium. Let PI and P2 be
pressures in the media. Then at each regular point P of the surface its
4.1 Visual Geometry of Soap Films. Minimal Surfaces 257

mean curvature H(P) is equal to ",(PI - P2), where", is a constant inde-


pendent of the point, and PI - P2 is the pressure difference in adjoining
media. In particular, the mean curvature of the interface is. constant (at
all regular points).

The constant", has an important physical meaning. The number 1/ '"


is called the surface tension coefficient (on the interface); it characterizes Surface tension
the properties of the media. Thus, the mean curvature of the interface is coefficient
constant. Consider two cases:
1) the mean curvature is equal to zero (this takes place if and only Zero mean CUNature
if the pressures in the neighbouring media are equal); +-+ minimal surfaces
2) the mean curvature is nonzero; in this case it is strictly positive Nonzero mean CUNature
(sign reversal can be achieved by reversing the normal to the· surface). +-+ soap bubbles

Both cases are modelled well in the physical experiment familiar to Physical experiments
every reder. This is the experiment with soap films. We take an aqueous
soap solution, plunge a closed wire contour into it and take it off. Then Wire contours
on the contour there forms an iridescent soap film. It can be interpreted and soap films
as a boundary between two adjoining gas media. Indeed, on each side
of the film there is the air with equal pressures PI and P2 since the film
is non-closed, and therefore far from the film both gas media adjoin and
interpenetrate, diffuse into each other. Thus, we have realized case 1 (see
above).
Case 2 is realized as follows. Into a soap solution we plunge a thin
tube and then blow through it a soap bubble. It will separate from the
tube and fall down smoothly, acquiring the shape of a sphere. We have
here two media: the interior. volume of the air with the pressure PI,
and the exterior volume of the air with the pressure P2. It is clear that
PI > P2 and that the system is in equilibrium only due to the fact that the
surface tension forces on the soap film compensate the excess of internal Soap films
pressure as compared with external. Thus, the mean curvature is here with boundary =
strictly positive (and constant). minimal surfaces
Let us concentrate on the first case, i.e. on soap films with boundary.
The mean curvature of such films is zero.
This class of surfaces can be shown to admit another, equivalent
description. Namely, the mean curvature of a two-dimensional surface is
equal to zero if and only if it is locally minimal, i. e. any sufficiently small
258 4. Visual Images in Some Other Fields of Geometry

perturbation of the surface (concentrated in a small neighbourhood of


an arbitrary poinlj does not diminish its area. On this ground, soap films
of the first type (i.e. those possessing zero mean curvature) are called
minimal surfaces. Thus, an equilibrium boundary between two physical
media with equal pressures is always a minimal surface.
Locally minimal 2-surfaces Since the theory of minimal surfaces is very closely connected with
are exactly the surfaces boundaries between physical media, with optimal shapes of membranes,
of zero mean curvature etc., it is in the centre of constant attention of a great number od modem
researches originating from scientific works of the XVII~XVIIIth cen-
turies. For the review of some of these trends see Refs. [33], [29]. These
investigations are one of the most important branches of variational cal-
culus [77], [78].
Here we shall only demonstrate some visual properties of minimal
surfaces underlying profound mathematical studies.
What is the geometrical meaning of equality to zero of the mean
curvature? From the definition it immediately follows that at the point
where the mean curvature is zero both the principal curvatures are equal
in the abs9lute value and opposite in sign since AI + A2 = O. From the
definition of the numbers AI and A2 it is seen that locally, near the point
P, the surface is organized like a saddle (FigA.1.2). In one direction the
Fig.4.1.2
surface goes up (relative to the normal). This direction corresponds to
the positive principal curvature, for example, AI. In the other direction
(it can be proved that it is always orthogonal to the first one) the surface
goes down.·This direction corresponds to the negative principal curvature
A2. So, near the point of zero mean curvature the surface is organized
like a mountain pass, a saddle.
Minimal surfaces Thus, all the points of a minimal surface are saddle points. In a
have non-positive particular case, a saddle may degenerate into a flat surface. Clearly, the
Gauss curvature mean curvature of an ordinary plane is equal to zero. Here both principal
curvatures are zero.

4.1.2 Some Examples of Minimal Surfaces

Catenoid, helicoid, We shall give some classical examples of minimal surfaces which have
Schwarz surface been investigated by mathematicians and physicists (from different points
of view) beginning with Laplace, Lagrange, Plateau.
4.1 Visual Geometry of Soap Films. Minimal Surfaces 259

1) The catenoid (Fig. 4.1.3). It is formed by rotation around the


axis of the curve given by the graph y = a ch x/a, where a = const.
This curve is also referred to as a chain line with the shape of a sagging
heavy chain (in the vertical gravitational field) fixed at two points. The
-~
tJ)
-+-+-

boundary of the catenoid in Fig.4.1.3 consists of two coaxial circles Fig.4.1.3


(generally speaking, of distinct radii).
2) The helicoid (Fig.4.1.4). It is formed by uniform rotation of the
straight line I intersecting the vertical z-axis and going uniformly up, so
that each point of the straight line I draws a helical line.
3) The Schwarz surface (Fig. 4.1.5). Its boundary contour is formed
by four edges of a regular tetrahedron. Then the Schwarz surface is a
saddle surface bounded by four rectilinear segments.

The Schwarz surface is interesting not only by itself but also for the
fact that it permits construction of new minimal surfaces. It is relevant
to formulate here the principles met by minimal surfaces.
Property 1. If a minimal surface has a free boundary 'Y which is
allowed to slide freely along a fixed two-dimensional surface M in ~.3,
then the minimal surface intersects the surface M along the curve 'Y at
Fig.4.1.4
a right angle 90° to the surface.
Property 2. Three smooth pieces of minimal surfaces which form
together a stable minimal surface and intersect along a common smooth
curve will necessarily make with one another equal angles 271"/3 = 120°.
Only four such singular edges (along each of which three sheets of the
minimal surface intersect) can meet at one isolated singular point. The
angles between any two singular edges coming to this point are equal at
this point to 109°28'16".
f'V

Property 3. If some portion of the boundary of the minimal surface


M is contained in a certain straight line, then the reflection of the surface
M' relative to this line is also a minimal surface, and the union of M and
M' forms a smooth minimal surface without a break on the rectilinear
region of the boundary. Fig.4.1.5
Property 4. If the minimal surface M meets a plane at a right angle,
its mirror reflection M' relative to this plane is also a minimal surface Reflection of the
and the union of M and M' forms a smooth minimal surface. minimal surface
Applying Properties 3 and 4 to the Schwarz surface, we can construct Periodic minimal
new, so-called periodic minimal surfaces. This fact is now widely used surfaces
260 4. Visual Images in Some Other Fields of Geometry

in architecture. For example, the famous Olympic Stadium in Munich


(built in 1972) was designed using minimal surfaces. Its flexible roof
was raised on special rods and was somehow similar to the Schwarz
surface, i.e. proved to be optimal from the point of view of area and
stability.
Enneper surface 4) The Enneper surface is given in JR.3 by means of the radius vec-
tor a(r,8) = (r cos 8 - ~ cos 38, -r sin 8 - ~ sin 38, r2 cos 28). Try
to construct is graph.
Sherk surface 5) The Sherk surface is given by the equation 1/a In cos ay / (cos ax:
=Z, where a =const; or eza =cos ay / cos ax. Construct the graph.

4.1.3 Visual Material

Vortex-type motion The vortex motion frequently arising at the boundary between two me-
dia, when the system is not in equilibrium, is presented in Fig.4.1.6.
Shown here are vortices in the atmosphere arising near the earth surface.
In Fig.4.1,7 there are objects closely connected with the theory of sur-
Flow in capillaries face tension of liquid. These are capillaries filled with liquid and drops
separating from the moistened surface by gravity. The theory of flow
Theory of meniscus in capillaries, the theory of meniscus and drop shape constitute one of
Drop shape the most interesting fieldS of application of the concepts of differential
geometry [79]-[81]. Within the theory of minimal surfaces one can de-
Drop separation scribe not only the process of drop separation, but also the process of
transformation of an unstable cylindric liquid column into a set of indi-
vidual drops, i.e. stringing out on a common horiiontal thread - the axis
of the initial liquid .cylinder. Figure 4.1.8 shows the boundaries between
media, vortex-type motions in the atmosphere or in the boundary liquid
or gas layer.

Page 261, Fig. 4.1.6. Boundary between two media; Laplace-Poisson the-
orem. J.S. Bach: "Matthew's Passion"; the last talk with Jesus

Page 262, Fig. 4.1.7. Capillares and surface tension of fluid

Page 263, Fig. 4.1.8. Analytic functions and modelling of a vortex-type


motion in the liquid. Egyptian priests and military judges in the court
264 4. Visual Images in Some Other Fields of Geometry

4.2 Fractal Geometry and Homeomorphisms

4.2.1 Various Concepts of Dimension

In the preceding chapters we mainly dealt with symplicial complexes,


Concept of dimension polyhedra and smooth manifolds for which the concept of dimension
is intuitively clear and mathematically easily dermed in a natural way
because it reduces to the concept of dimension of a Euclidean space.
There are different We believe that the concept of dimension ofll~n is known from elemen-
concepts of dimensions tary geometry and linear algebra. The dimension of a manifold is the
dimension of Euclidean balls (domains) of which the manifold is glued.
The dimension of a polyhedrol!. is the maximum dimension of simplexes
entering in its composition. But in mathematics, mechanics and physics
some manifolds are encountered (and particularly often in recent years)
for which the concept of dimension needs special discussion and, more-
over, not one but several distinct dimensions can be defined for them in
a natural way. These dimensions may not coincide. It is intuitively clear
that we are speaking of sets whose local organization is "substantially
worse" than that of open regions in a Euclidean space. Strictly speaking,
For "nice spaces" different concepts of dimension can be defined for an arbitrary topologi-
al/ concepts of cal space. But for "good spaces", to which manifolds and finite polyhedra
dimensions coincide belong, all these numbers (dimensions) coincide. But as soon as we pro-
ceed to the consideration of more complicated exotic (and sometimes in
Strange attractors a sense "pathological") objects, different concepts of dimension lead us,
generally, to different numbers. It was believed earlier that this occurs
mainly for the class of spaces encountered seldom in practice, for in-
stance, in physics, etc. But it has recently become clear and is widely
known already that such anomalous (from the point of view of dimension)
objects are encountered in classical fields of mathematics connected with
concrete physical applications, for example, in the theory of differential
equations, dynamical systems (the so-called "strange attractors"), etc. In
this connection, the interest, in particular, in the analysis of different con-
Fractal geometry cepts of dimension has become lively again. The concepts of "fractal"
and "fractal geometry" have been formulated (see the well-known book
by Mandelbrot [10]). A whole complex of questions arising in the study
of "locally complicated" objects is investigated in the framework of this
4.2 Fractal Geometry and Homeomorphisms 265

scientific direction. We shall only briefly touch upon this rich topic. We
start our discussion with the concept of dimension.
Recall that the family of sets Ui of a topological space X is called
the covering (of the space X) if each point x from X belongs to a Covering of a space
certain set Ui. An open covering consists of open sets, while a closed Open, closed, and
covering consits of closed sets. We now consider only finite coverings. finite coverings
The multiplicity of a covering is the largest of the integers n, such that Multiplicity
there exist n elements of the covering (i.e. sets) Ui having a non-empty of a covering
intersection.
A family of sets Vj is called inscribed in the family of sets Uj
provided that each element of the family Vj is contained at least in one
elements of the family Uj.
We shall now formulate the concept of topological dimension which
goes back to the works by Brouwer, Lebesgue, Menger, Uryson. For
details see e.g. Refs. [10], [82]. For simplicity we consider the class of
compacts (compact sets), i.e. such that from among any of their open
coverings one can always choose a finite covering.

Definition. The topological dimension dim X of a compact X is the Topological dimension


smallest of the integers n, such that in any open covering of the space dim X
X one can inscribe a closed covering of multiplicity ~ n + 1.

If there are no such numbers n, the dimension is assumed to be


equal to infmity. The visual meaning of this definition is fairly simple.
For example, when n = 2 it states that any two-dimensional "ground"
(compact) can be tiled with arbitrarily small stones (closed sets), so that
the stones adjoin one another not more than by three. At the same time,
this ground cannot be tiled by arbitrarily small stones so that they adjoin
only by two. Therefore, the statement that, for example, a square has two
dimensions is sometimes called "The theorem on tilings" when streets The theorem on tilings
or squares are paved with cobblestone, stones always adjoin by three,
and one can avoid their adjoining by four. In case a three-dimensional
volume is filled with sufficiently small stones (e.g. brick-work in a large
cavity) there necessarily arise adjoinings by four.
A great achievement of mathematical thought was the discovery of
the fact that this number of adjoinings (the multiplicity of coverings)
naturally contains the concept of dimension. This contribution is due to
266 4. Visual Images in Some Other Fields of Geometry

Lebesgue (1911). Figure 4.2.1 shows a small covering of multiplicity one


of a one-dimensional segment by closed sets. It also shows the covering
of multiplicity two of a two-dimensional square. It is characterized by the
fact that to each knot (vertex) there comes exactly three one-dimensional
edges whose union is the boundary of the elements of covering.

Spherical Another approach to the concept of dimension is the idea expressed


Hausdorff measure by Hausdorff (1919) and .later developed by Bezikovich. This idea was, in
fact, first expressed by Caratheodory (1914). To formulate this concept
I I I , I I I
of dimension, we need the spherical Hausdorff measure. Let X be a
compact subset in a metric space, for instance, in a Riemannian manifold
or a Euclidean space. Consider a finite covering of this compact by m-
dimensional balls Dr(ei) with radii ei. Let "1m be an m-dimensional
volume of a standard m-dimensional unit ball (in a Euclidean space
]Rm). The number "1m can be calculated explicitly, but we do not need
this formula here, and therefore it is omitted. Then the volume of the ball
of radius ei will be written as "Imer. From now on we do not assume
Fig. 4.2.1 the number m to be an integer. Let m ~ 0 be an arbitrary non-negative
real number. Calculate the sum L:i 'Ymer. Taking another covering of
the compact X with balls, we shall obtain another value of this sum.
Consider all possible coverings of the compact X with balls of radii not
exceeding a fixed number {! and calculate hI! = inf!i<1! L:i 'Ymer, where
inf is taken over all such coverings. If {! tends to zero, this inf may only
increase. Therefore, there always exists a limit lim(!->o hI!' This number
is always non-negative (it can be either finite or infinite) and is called
Hausdorff dimension an m-dimensional spherical Hausdorff measure of the set X. We shall
denote it by JLm(X). We now proceed to the1:oncept of dimension. Note
once again that in the definition of the Hausdorff measure the number
m was not at all assumed to be an integer.
Let X be a compact in a metric space. Bezikovich has shown that for
each X there always exists a real number D, such that the m-dimensional
Hausdorff measure of the compact X is infinite when m < D and on
the contrary equals zero when m > D.

D(X) ~ dim X for any X Definition. The number D is called the Hausdorff dimension of the
compact X (or the Hausdorff-Bezikovich dimension).
4.2 Fractal Geometry and Homeomorphisms 267

In the indicated sense the number D = D(X) is the critical dimen-


sion. As distinct from the topological dimension dim X, the dimension

- - tt - -
D(X) need not necessarily be an integer. In particular, these two dimen-
sions may not coincide. They are related by the inequality D ;::: dim, i.e.
D(X) ;::: dim X for any X. For "good spaces" these dimensions coin-
cide. At the same time there are examples of complex spaces for which •• • • •• ••
D>dim. t

Definition. (see Ref. [10]). The fractal is a set for which D > dim. Definition of (ractal

In particular, any set for which the dimension is not an integer is a


fractal. For details see the book by Mandelbrot [10].

4.2.2 Fractals

Consider the so-called Cantor set(called so after George Cantor). Its


simplest version is constructed as follows. Consider a unit segment from
oto 1 on a real axis. Divide it into three equal parts after which discard
the middle third, i.e. the open inverval from 1/3 to 2/3 (Fig.4.2.2). Each
of the remaining two thirds also divide into three equal parts, after which
discard their middle (open) parts. Repeating this procedure infinitely,
we obtain a closed set K on the unit segment. It is intuitively clear
(and can be proved exactly) that dim K = O. At the same time, the
Hausdorff dimension of the Cantor set is strictly positive and equal to
D(K) = log 2/log 3 : : : 0.6309. The construction of the Cantor set can
be transformed in such a way (we leave it to the reader to guess in Fig. 4.2.2
what particular way) that its Hausdorff dimension will be equal to any
aforehand given number >. in the interval between 0 and 1. The equality
dim K = 0 will hold as before. Thus, the Cantor set is a fractal. An Cantor set
example of the fractal is also the so-called Serpinsky carpet - a closed Serpinsky carpet
set on a Euclidean plane. Its construction is pratically similar to that
of the Cantor set, but one should start with a flat square, step by step
discarding from it increasingly small squares.
Fractals naturally appear also in the theory of complex transforma-
tions of a plane. On a plane we introduce a complex variable z = x +iy
and consider a complex polynomial fez) of degree k. It can be inter-
268 4. Visual Images in Some Other Fields of Geometry

preted as a complex transformation of the plane, as a map f: C - C


given by the formula w = f(z). This map can be iterated by considering
the infinite sequence of its powers fn = r, where fn(z) = f (fn-I(Z)) =
f(f( . .. f(z)) . .. ) (in times). For example, fl (z) = f(z), h(z) = f(f(z)),
Fractals in the theory of etc. Let z be an arbitrary point of the plane. Applying to it the inifinite se-
complex transformations quence of transformations r, we obtain the inifinite sequence of points
of a plane z, f(z), f2(z), ... , r(z) . .... In other words, we consider the orbit of
the point z under the action of iterations of the map f. It is convenient
to consider a complex compactified plane as a two-dimensional sphere.
To this end one should add an infinitely remote point to a Euclidean
plane. Then the infinite sequence of points r(z) necessarily has limit-
ing points on the sphere. We say that a subset Q on the sphere (or on
the plane) is invariant under the map f if it is carried into itself under
the action of all iterations r, n = 1,2, .. . . Such sets often appear to
be fractals [10]. Such objects appeared for the first time in well-known
Fatou, Julia, works by P. Fatou (1878-1929) and G. Julia (1893-1978). Since f is a
Mandelbrot complex polynomial, the mapw = f(z) always has fIXed points. They
are defmed as roots of the equations f(z) = z, i.e. f(z) - z = O. If the
degree of the polynomial f is equal to k, this equation has k roots. Some
of them may coincide, i.e. to be multiple roots. The roots are represented
by points Zl, .•• , Zk on the compactified plane C = ]R2 U00. Suppose at
some of these points Zi there holds the inequality Id/dz f(zi)1 < 1. Then
the map f is "contracting" at this point, i.e. the root Zi is the point of
attraction of some other points of the plane. Next we may consider on
Domains of attraction the plane the domain of attraction of the root Zi, i.e. the set of points
Z for which limn-too r(z) = Zi. Such sets also often appear to be frac-
tals. They may have an exceptionally complicat~d and at the same time
beautiful form [10]. We may also consider the following set of points
{z : limn-too I r(z) 1= oo}, i.e. the domain of attraction of an infinitely
remote point. These sets are also frequently referred to as fractals.

4.2.3 Homeomorphisms

Homeomorphism is one of the basic concepts in topology. Homeomor-


phism, along with the whole topology, is in a sence the basis of spatial
Fig. 4.2.3 preception. When we look at an object, we see, say, a telephone receiver
4.2 Fractal Geometry and Homeomorphisms 269

or a ring-shaped roll and first of all pay attention to the geometrical shape
(although we do not concentrate on it specially) - an oblong figure thick-
ened at the ends or a round rim with a large hole in the middle. Even if
we deliberately concentrate on the shape of the object and forget about
its practical application, we do not yet "see" the essence of the shape.
The point is that oblongness, roundness, etc. are metric properties of the
object. The topology of the form lies "beyond them". It is of interest that
we do not need any special training to learn to perceive this essence of
the shape because it is an inbom ability of our sight to see holes and
connections, regions and lines, adjoinings and intersections, cIosedness,
etc. We see all these visual properties since we are born, when the for-
mal thinking is not yet formed, and therefore it is difficult to pay special
attention to the properties of the object laid in our perception. It is per-
haps just for this reason that the science investigating these properties,
i.e. topology, developed later than the other fields of geometry, in fact,
only in the XX-th century.
In mathematics these "basic properties" are defined as invariants Invariants of
oj homeomorphism. Homeomorphism is a superposition of some figures homeomorphism
upon others, but a more free superposition than a rigid matching of
triangles in Greek geometry, in the proof of their equality (or inequality).
It is useful to imagine figures or bodies made of caoutchouc, but much
more elestic than it actually is. This allows us to expand, contract, twist
the object. Only breaks and glueings forbidden. For example, if we melt a
roll made of wax, this will be an unlawful transformation because at some
moment glueings and breaks will appear. Although the transformations
admitted by homeomorphism can be stubstantial, something still remains
unchanged. This "something" is just the subject matter of topology. For
example, this is the number of connected pieces or the number of holes.
The two problems often to be solved in topology are how to establish
that two figures are homeomorphic (if at all) or how to show that they
are not homeomorphic. To solve the first poblem, one should point out
a more or less explicit geometric construction (most often to present
a homeomorphism); to solve the second one, one should indicate the
property which is distinct for these ojects, i.e. the invariant which is
more convenient to represent in an algebraically calculate form.
In classical Greek geometry there exist two kinds of equalities. Two
triangles can be matches either by moving them continuously along the
270 4. Visual Images in Some Other Fields of Geometry

IIIIIII I-+-u,m-u plane or (if they are axissymmetric) only by taking them out of the plane
and changing their position in space. In the latter case this is the matter
Fig. 4.2.4 of orientation. More complicated cases are possible too. Take a "comb"
(Fig. 4.2.4) and bend aside one of its teeth. Doing so, assume the comb to
be a one-dimensional curve. The figures obtained are, of coUrse, home-
omorphic. But the establish the homeomorphism, one should not pay
attention to the enveloping plane. Neither of them can be mapped into
another via homeomorphism (prove itl). When considering the home-
omorphism problem as depending on the position of the figure in the
Problem of embedding enveloping space, we are led to the problem of embedding fonnulated as
follows.
Let in ]Rm (it is only for the sake of simplicity that we take a
Setofa/l Euclidean space) two figures A and B be given of which we know that
self-homeomorphisms ~eyare homeomorphic. Is the possible to construct a homeomorphism
of Euclidean space h: 1 m - t ]Rm such that h(A) = B?
In a specified fonn, we are given a homeomorphism y: A - t B and
it is required that h be so constructed that for points x from A we have
h(x) = y(x), i.e. that h might continue y from A to the whole of]Rm.
It would seen reasonable that before turning to these general prob-
lems, .we should first investigate the set of all self-homeomorphisms of
Klein's the space ]Rm. This would correspond to the well-known F. Klein's Er-
Erlangen program langen program - the geometry is created by the transformation group for
Group of homeomorphisms ]Rm; in this case this is the group of homeomorphisms as the widest of
all groups of continuous transformations. But investigation of the group
of homeomorphisms of]Rm (denoted here by H(]Rm)) has been delayed
up to recent decades and is not yet quite completed now. We shall briefly
touch upon some of the results obtained approximately after 1960.
We shall begin with one construction proposed by an outstanding
topologist of the XX-th century 1. Alexander. It is extremely simple. Con-
sider homeomorphisms motionless on a ball D;' of radius p and centre
at O. This means that h(x) = x if x lies in D;'. It turns out that for such h
one can easily construct a one-parameter family of homeomorphisms or
an isotopy to join h with an identity homeomorphism e (where e(x) = x
for all x fonn ]Rm), i.e. such a family ht, where t varies fonn 0 to I and
hi = e, ho = h and ht(x) depends continuously on the arguments x and
t. Moreover, each ht is a homeomorphism. To prove this, we consider
the family of homotheties kt : ]Rm - t ]Rm, i.e. similarity transformations.
4.2 Fractal Geometry and Homeomorphisms 271

(expansions, contractions) of the space ]Rm with coefficients t: x --? tx.


Then it suffices to put ht = kt I hkt for t f 0 and separately ho = e.
This construction admits many modifications. For example, the sub-
stitution c(x) = x/lxl2 (inversion; 0 is discarded) readily gives a similar
result for homeomorphisms identical outside the ball Dm , or simple for
homeomorphisms of Dm identical on its boundary. As a result, we can
obtain a homeomorphic correspondence between the space of homeo-
morphisms preserving direction at the origin and the space of homeo-
morphisms preserving direction at infinity. The latter condition can be
defined in essentially different ways, and we invite the reader to decide
what the needed defmition should look like. This contribution is due to
1. Kister.
For direction-preserving homeomorphisms we incidentally obtain a
technical result: each of them is representable as a composition of two
homeomorphisms each of which is identical on some ball. The condi-
tion is stronger (from the formal point of view) than the possibility of
constructing an isotopy joining the homeomorphism with the identity
transformation. Therefore, for a long time attempts have been made to
show that each orientation-preserving homeomorphism (this condition is,
of course, necessary) is representable as a finite composition (which can
always be reduced to two) of homeomorphisms identical each on its own
ball. Such homeomorphisms are called stable and the problem itself the
stability of homeomorphisms. It is now solved in all dimensions except Stable homeomorphism
four, and its solution appeared to be linked with the most fundamental
questions of the topology of manifolds, the answers to which have been
found in recent decades.
The next step made by R. Kirby [83] in 1968 looks like a rather
on
simple extension of the methods based the Alexander's remark (see
above). But it is precisely this remark that has ultimately led to the
discovery of connection of different results obtained before in isolated
fields of the topology of manifolds and, in the end, to the solution of the
indicated problems.
Consider the homeomorphisms of a torus Tm. Represent it as a quo- General homeomorphisms
tient manifold by the action of]Rm of a group of integer-valued linear of a torus
transformations corresponding to translations into linear combinations of
m independent vectors which form an orthogonal frame. Let p be a nat-
ural projection of the space]Rm onto Tm. It maps into one point a whole
272 4. Visual Images in Some Other Fields of Geometry

orbit of this group, i.e. the set x + L:k Ckek (where Ck are integers). To
each homeomorphism h of the torus there correspond homeomorphisms
h of the space ~m which cover this space. As h(x), where x is a point
from ~m, it is natural to take one of the points lying over the point
h(P(x», and it suffices to make the choice for one of them because for
the rest it is uniquely determined from the continuity condition. So, we
have the condition ph(x) = hp(x) which can be written as the commuta-
~m~~m
tivity condition for the diagram p1 p1
Tm~ T m

We obtain the map HT(~m) - t G(Tm), where HT(~m) denotes


Stable homeomorphisms the group of homeomorphisms mapping each set of points lying over
of the torus one and the same point of the torus (the orbit) into another similar set
over another point of the torus. In this case the homeomorphisms can
be said to be permutative with the action of the indicated group. Each
homeomorphism h of the torus generates a linear transformation A with
an integer-valued matrix. It is most easy to define A as the transformation
of the fundamental group of the torus induced by homeomorphism. An
extensive definition will be following. Let the point 0 be motionless under
the action of covering homeomorphism. This condition does not violate
Covering homeomorphism generality. Then this covering ~omeomorphism preserves the whole grid.
(lattice) of points L: Ckek with integer coefficients Ck and determines on
it the required linear transformation.
Consider torus homeomorphisms generating an identity transforma-
tion (an identity matrix) in ~m. The group of these homeomorphisms
will be denoted by Ho(Tm). Any homeomorphism can be transformed to
Ho through the transformation h - t z-l h, where z is a standard home-
omorphism generating the same linear transformation as h. Homeomor-
phisms from Ho are direction-preserving at infinity since they displace
each point not more than by the size of the fundamental region (in our
case, by the diameter of a unit cube). In particular, they are stable. Since
the stability is obviously local, the corresponding homeomorphism of the
torus, which we have taken in fact arbitrarily, is stable too. The arbitrari-
ness was restricted by the fact that the homeomorphism was taken from
Ho(rm), but any homeomorphism of a torus is a composition of such a
4.2 Fractal Geometry and Homeomorphisms 273

homeomorphism and a standard one. A standard homeomorphism is, of


course, stable if it is orientation-preserving (and this we have assumed).
Thus, the important step made by Kirby appeared to be possible
because he overcame the inertia of thinking, and in the obviously lo-
cal problem, which therefore obviously required consideration only of a
local model of manifolds, i.e. a Euclidean space, he considered a man-
ifold different from ~m and from the sphere (to which the attention of
researchers was mostly given at that time), namely, the torus.
The homeomorphisms of the torus all appeared to be stable without
exception. To make further use of this remark, Kirby has to apply the
technique to be described below.
We shall first mention one consideration, also a remote descendant
of Alexander's construction, with the help of which the isotopy joining a
homeomorphism with the identity homeomorphism is proved to exist for
any homeomorphism sufficiently close to the identity one. This isotopy
may actually be constructed (see below) as continuously depending on
the homeomorphism. In the topological language this property is called
the local contractibility of the homeomorphism group. This theorem, for- Local contractibility of the
mulated by A.V. Chemavsky, is one of the basic theorems in the above- homeomorphism group
mentioned development in the topology of groups of homeomorphisms
of manifolds [84].
The crucial point of the consideration can be formulated as a lemma.
It consists in the fact that a small topological displacement of a cube in
~k can be covered by the topological perturbation of the whole of~k.

Lemma. Let Iff be a cube with side 2p and centre at the origin. There
exists c > 0, such that for each homeomorphism (embedding) g: If-+
~k displacing points not less than by c (i.e. for all x we have: the
distance from x to g(x) is less than c) one can construct the isotopy
ht : ~k -+ ~k which joins on If map 9 with the identity map, i.e.
ho = 1, hi = 9 on If. Given this, ht depends continuously on 9 and is
identical for the identity embedding g.

The latter condition, in particular, means that the less the map 9
displaces the points, the closer the unknown isotopy is to identity. The
required covering is constructed only for the cube If, and the embedding
274 4. Visual Images in Some Other Fields of Geometry

is determined on the larger cube If. This is necessary for the proof and
is no obstacle for using the lemma. See Ref. [84].
The general result reported by Chernavsky [84] consists (if formu-
lated very briefly) in local contractibility of the homeomorphism group
of any manifold. For a noncompact manifold one should specify the
topology of homeomorphism space, but we do not discuss it here.
As a simple application of this result, the reader may answer the
question which was formulated by D.V. Anosov and gave impetus to the
proof of the theorem.
Suppose a homeomorphism, which differs little from the identity
one, is given on the boundary of a manifold. Can it be continued to the
homeomorphism of the whole manifold?
Returning to stable homeomorphisms we note that the above-men-
Stability of tioned results make it possible to establish that homeomorphisms differing
torus homeomorphisms little form the identity one are stable. Furthermore, this consideration has
not dimensional restrictions, which isa pleasant exception to the basic
results of the modem topology of manifolds. The arguments needed to
translate the Kriby's proof of stability of torus homeomorphisms into the
proof of stability of homeomorphisms ofl.m run into rather complicated
facts of the topology of manifolds proved only for m f 4, and therefore
the case 1.4 requires special consideration.
It should be noted that from the torus one can easily go over to the
manifold T x I., where T denotes the torus Tk-I. It turns out that each
homeomorphism M x I., where M is a compact manifold (or generally
a compact), is isotopic to the homeomorphism covering the homeomor-
phism M x 8 1•
The concept of a stable homeomorphism admits an extension which
we introduce without claiming a complete generality. We call a home-
omorphism of a region in I.k onto a region in I.k stable if it becomes
linear in the neighbourhood of a certain point by means of a composition
with a homeomorphism identical outside some neighbourhood of the im-
age of this point. This may be any arbitrary point of the given region.
This means locality of the stability.
Let now T x I. be represented as a subset (subregion) in ]Rk. This is
obviously possible for k = 2, and for larger k is obtained by a successive
Fig. 4.2.5 construction of "surfaces of rotation" (Fig. 4.2.5). Suppose we are given
a homeomorphism h: I.k -+ I.k • What can be said about h(T x I.)? This
4.2 Fractal Geometry and Homeomorphisms 275

is an open subset in I.k and is, therefore, a manifold, a triangulatable one


(as any open subset in I.k ). In the technical language of the topology of
manifolds one can say that a piecewise linear structure of this manifold
is defined.
At this point the problem of homeomorphisms ofl.k first meets with
the problem of structures on manifolds, which is one of the basic prob-
lems in modem totopolgy of manifolds. The contiguity of these problems
turned out to be progress made in the development of the topology of
manifolds since the 1970s.
The piecewise linear structure is understood as a complete set of Piecewise
triangulations of a given manifold, all of them being connected with one linear structure
another so that each two have a common subdivion. Moreover,the tri-
angulations themselves should be locally isomorphic (perhaps, after the
subdivion) to triangulations of a region in I.k , which is of course fulfilled
in our case. The fundamental problem conventionally called Hauptvermu-
tung consists in the statement of uniqueness of such a structure. Namely,
ifa homeomorphism is established between two manifolds with piecewise
linear structures, then between them one can also establish a piecewise
linear homeomorphism, i.e. a homeomorphism linear on each simplex
of a certain subdivision. If this problem were solved at least for T x I.
(said Kirby in 1969), everything would be all right: we would have a
homeomorphism h: T x I. -+ h(t x I.) linear on simplexes, and since
the homeomorphism h-1h is stable (as a homeomorphism ofT x I.), so Problem of
will be h at least on T x I.; by virute of locality of this property, h will stable homeomorphisms
be stable everywhere.
The question of uniqueness of structures on T x lR. appeared to be
rather mature by that time. By the way, it soon became clear that this
question is generally solved negatively, but Siebenmann then made an
important remark. The point is that there exists an obstacle (a homology
class with coefficients in the group Z2) responsible for non-uniqueness
of the structures. It was unknown at that time whether the required class
is equal to zero. But it was known that it becomes zero in passing over
to some finite-sheeted covering which is again homeomorphic to T x I..
By virtue of locality of our problem, it is sufficient that we establish
stability of the homeomorphism covering the given one.
This solution of the problem of stable homeomorphisms (for k 4) r
would in itself be a rather peripheral fact of the topology of manifolds
276 4. Visual Images in Some Other Fields of Geometry

and would be of interest mainly as an example of establishing an unex-


pected contiguity between two branches of the study of manifolds. In the
course of this, however, profound properties of the various structures on
manifolds were discovered. The discussion of different visual properties
of homeomorphisms are contained also, in particular, in [86]-[90].

4.2.4 Visual Material

Fractals Fractals appear naturally in modelling the Brownian motion, in con-


Random processes structing sets resembling the Cantor set, in studying random processes.
Random sequences Examples of random sequences of numbers are decimal decompositions,
for example, of numbers 7r and e. Figure 4.2.6 demonstrates typical pic-
Complex dynamiCS tures arising in fractal geometry in the study of the complex dynamics
(see Ref. [10]) of complex transformation of a plane. Fractals arise in
dynamic systems in the study of translations along the trajectories of the
system. It is often useful to imagine the structure of fundamental regions
of a given action of the group. Various elements of fractal geometry are
"From chaos to order" presented in Figs. 4.2.7 and 8. Figure 4.2.7 (from the cycle "From chaos
to order"), which has something in common with the ideas expressed by
Shiryaev in his book [16], provides an abstract interpretation of the idea
of the probability theory and statistical fractal geometry concerning a
gradual discovery of harmony and order in a primarily seeming chaos of
images. Figure 4.2.8 is an attempt to illustrate the idea of mathematical
(geometrical) infinity closely connected with the idea of fractals. Infinite
variations and the variety of one and the same habitual image resem-
ble distortion of fundamental region under discrete action of an infinite
group. Figures 4.2.9,10 and 11 are devoted to the attempt of modem
geometrical interpretation of some scientific ideas which excited already
Medieval medieval scholars and painters (in particular, Bosch, Bruegel, Diirer and
scientific ideas others). All of them were interested in the geometrical theory of per-
spective, developed the principles of a correct drawing of objects, were
pioneers in discovering mathematical mechanisms of optical illusions,
etc. The creative activity of authors in those times was naturally under
the influence of a specific (and very interesting) scientific world outlook
which also found its expression in drawings and paintings of that epoch.
Such is, for example, the well-known Bruegel's engraving "Alchemists".

Page 277, Fig. 4.2.6. Complex dynamics and transformation of the plane.
The last battle with titans; ancient and medieval Greece
278 4. Visual Images in Some Other Fields of Geometry

Figure 4.2.9 can be interpreted as a mathematical variation of this theme


whose essence is visual modelling of the idea of mathematical infinity
which was the subject of lively interest and excitation for many scholars
beginning from Renaissance. In this sense Figs. 4.2.9, 10 and 11 are an
attempt to imagine how medieval geometricians would have depicted the
Medieval and modern ideas, legends and other things that stirred their minds if they had com-
concepts of mathematical manded the concept of mathematical infinity to the extent accessible to a
infinity mathematician nowadays. Of course, many scielltific concepts of authors
of the Middle Ages have now lost their topicality and seem rather strange
to the contemporary reader and viewer, but many modem geometrical
ideas have naturally grown from the soil fertilized by them.
Therefore, we think it interesting to synthesize in several graphi-
cal works somewhat strange medieval images and modem mathematical
concepts of infinity, deformation, continuity. Figures 4.2.9,10 and 11
present both many symbols and images of medieval science (geomet-
rical and astronomical devices, chemical materials and equipment, etc.)
and purely modem objects (optical lenses, apparatus, etc.). But this mix-
ture is not mechanical, it is hereditary because it is viewed by the eyes
of today' s spectator who feels the relation between thorough mathemat-
ical investigations of our time developed for example from the attempts
to prove the Fermat theorem (see the fragment in Fig.4.2.l1) and the
peculiar scientific thinking which populated the world of that time with
strange fantastic images.

Page 279,. Fig. 4.2.7. Idea of mathematical infinity and fractal dimension.
From cycle: "From chaos to order". Fanatics

Page 280, Fig.4.2.8. Idea of mathematical infinity. Prediction of a future


by ancient god

Page 281, Fig.4.2.9. Comparison of medieval and modern concepts of


mathematical infinity. Variation of the theme of Bruegel's "Alchimists".
From the cycle "Conversation with medieval authors"

Page 282, Fig. 4.2.1 0. Modern mathematical variation on the theme of


well-known biblical book "Revelation" (St. John); from the cycle "Conver-
sation with medieval authors"

Page 283, Fig.4.2.11. Modern mathematical variation on the theme of


well-known medieval legend "Temptation of Saint Anthony"; from the cy-
cle "Conversation with medieval authors"
284 4. Visual Images in Some Other Fields of Geometry

Visualization 4.3 Visual Computer Geometry in the Number Theory


and computers
in number theory At first glance the theory of numbers is deprived of any geometricity.
But this is actually not the case. At the contemporary stage of develop-
ment of computers it has become possible to explain to a wide range of
readers that visual geometry helps not only to illustrate some abstract
situations from the number theory, but sometimes also to solve new
Visual problems. Visual computer geometry is of help in finding a mathemati-
computer geometry cal statement, after which there follows the stage of its logical proof. In
such a sophisticated theory as number theory it is sometimes very diffi-
cult even to formulate a plausible hypothesis; Modern computers permit
a correct choice of hypothesis after processing many experimental data
and representing them in the graphic form. Much depends on the way of
representation. The information should be aptly coded by visual images
so that the pictures appearing in the computer display might help the
researcher to fine the correct direction of subsequent steps in his study.
We now consider the problem of representation of natural numbers
n :::: 1 as the sums

n =ni +n2+" '+n~ , (1)


where all the numbers 1ti, 1 ::; i ::; s are non-negative integers.
Let us fix arbitrary values r ~ 2 and s ~ 1. In this case, all natural
numbers are divided into two classes. One class involves all natural
numbers representable in the form (1) and the second class involves
those which cannot be represented in the form (1) with given parameters
Waring problem r and s. There appears the general problem (the generalized Waring
problem): hwo shall we describe each of the indicated classes?
This problem is exceedingly complicated. We shall briefly describe
both the results obtained earlier and the recent ones. Paricular attention
will be given here to the results initially "groped for" using computer
drawings (after which they were proved mathematically).
This work was done by A.A. Zenkin [91], [92]. We shall describe
some of these results.
We agree that the natural numbers not representable in the form
(1) will be black while those representable will be white. In the origi-
nal experiments [91], [92] the colours were more spectacular, they were
4.3 Visual Computer Geometry in the Number Theory 285

respectively red and green. Now consider one of the possible ways to
present a natural series on a computer display.
Take an infinite strip and mark it into equal squares in which write 1112.131416' 16 1; I
natural numbers successsively (Fig.4.3.l). Fix an integer d and divide Fig. 4.3.1
this infmite strip into pieces of length d. This number will be called the
image modulus. Recall how the picture is formed on a television screen.
An electron beam runs through the first horizontal line, jumps over to the
beginning of the second line, runs it through, etc. As a result, the beam
sweeps the whole of the square screen creating a picture. We shall do
the same. We lay the first segment of length d onto the first line of the
display, the second segment onto the second line, etc. Then the display
gets filled with natural numbers form I to a certain N determined by the
display size.
Since we are primarily interested in the property of natural numbers
to be "representable or non-representable" (in the form (I)), we do not
need the absolute value of the numbers. If needed, this information can
be restored by counting the number of squares on the display. So, mark-
ing non-representable numbers in black and representable in white, we
see on the computer display a spotty black-and-white carpet consisting
of black and white squares (for given r, s, d). This carpet can be exceed-
ingly sophisticated. Since our basic problem is a visual determination of
regularities in the spotty carpet of numbers, its solution can sometimes be Visual determination
sped up by adding sound information. The display of each black square of regularities in the
was accompanied by the sound signal "C" and that of each white square "carpet of numbers"
by the signal "G". As a result, the computer not only drew a certain set,
but also played ''music''. Its rhythmics is evidently some invariant of the
represented set and (what is important) does not depend on the number d.
Fix r = 2 and start increasing the parameter s = 1,2,3, .... Fix
the value of d. So, we are studying the question of representability of Any natural number
numbers in the form a) of the square of a certain number, b) of the sum is the sum of four
of two squares, c) of the sum of three squares, etc. squares (Lagrange)
Let us look at the character of picture variation on the display
(Fig. 4.3.2). At first (for s = I) almost the whole of the display is black.
Here and there one can see white squares. One can also see how few num-
bers are squares. As s increases, the display starts "whitening". Finally,
for s = 4 the whole display flares white. Black squares have vanished. For
286 4. Visual Images in Some Other Fields of Geometry

s = 5 the picture no longer changes: the display remains white. Clearly,


we have "seen" a certain regularity.

Computer hypothesis 1. The sum offour squares (i.e. s = 4) suffices to


represent any natural number in the form (1).

Visualization of It turns out that we have seen on the display the well-known La-
Lagrange theorem grange theorem which he proved in 1740.

The Lagrange theorem. Any natural number is representable in the


form of the sum offour squares.

Every mathematician knows and can give many examples from his
scientific work when it appears much more difficult to feel or "see" a
correct hypothesis than later to prove it. Visual images are particUlarly
often used in geometry and topology where one has to work with mul-
tidimensional objects which, in principle, do not always admit picturing
in a three-dimensional space.
By Fig. 4.3.2 the computer informed us about something more. For
s = 2 the carpet surely has a simple structure: we clearly see six vertical
black columns, i.e. columns entirely consisting of black squares. This is
also a well-known theorem.

The Euler theorem (1749). Let a natural number n have the form n ==
3,6, 7(mod8). Then all numbers of the form n·4 k , where k = 0, 1,2, ... ,
are not representable in the form (1), i.e. n· 4k f nI + n~.

We haven chosen mod d so well that the numbers indicated in the


Euler theorem which are not representable in the form (1) are positioned
exactly one under another on the display to form the black columns.
For s = 3, there remain not many black squares in Fig.4.3.2. The re-
maining white field is clearly seen to be organized into several oblique
strips concentrating along white segments and having the same slope an-
gle. Clearly, this is again not accidentally. But for s = 3, the carpet in
Fig.4.3.2 is not yet well pronounced. Let us try to change slightly the
picture so as to make it more "completed". We shall change modd. The
result is shown in Fig. 4.3.3, where we can see pictures for d = 22,24. It
4.3 Visual Computer Geometry in the Number Theory 287

[ 0 r

Euler theorem
!=2. (L.ELlfe'l.)

r n n

Gauss theorem
Lagrange theorem
S::: 3 (Gauss) Fig. 4.3.2

is quite clear that when d = 24, the picture acquires its "final" form. The White-black carpets
columns have become vertical. Of course, there are black squares out- of numbers
side columns, but we are now primarily interested in those black squares
which constitute the vertical columns. This is a no less famous Gauss
theorem.

The Gauss theorem (IS01). Natural numbers are not representable as


the sum of three squares if and only if they have the form n = Sk +7, k =
288 4. Visual Images in Some Other Fields of Geometry
I" III "II
I •• J' I ,
I, JI. •
0,1,2, ... , (i.e. n == 7(mod8)) and if they are obtained from these ones
, '-- 'IIIIIL
-I'
via mulitplication by any power of the numeral 4, i.e. if they have the
• I '-',

J'.. II,,;'".
'.JI.
form n· 41, where l = 1,2,3, ....
JI' " -. ...
..... .
JlI.
'."" I

"
" JI, ,__ All details of the Gauss theorem are clearly seen on the display

.
IIIIIL 1,'-
" -1", JI • (Fig.4.3.3). Note that in any k-th column on the display (modd) there
.,.,
...J' •• '.'",
J' '.'- I,
stand numbers of the form n == k(mod d), i.e. numbers which when
divided by d give one and the same residue k. The main series of columns
" . JI •• (continuously black) are visually distinguished in Fig.4.3.3. Therefore,
". JI-. '..
mod 22.
they consist of numbers which when divided by 8 give a residue equal
to 7 (numbers == 7(mod8)), i.e. numbers of the form n = 8k + 7, k =
0,1,2, .... Employing the sweepingout procedure, which in any set of
numbers finds all subsequences of the type of arithmetic and geometric
progresssions, one can readily make sure that for d = 24 all the rest of
the black squares of the picture in Fig. 4.3.3 are elements of geometric
progressions of the for n· 41,1 = 1,2,3, ... , where n is the element of
one of the columns of the principal series.
So, the intrinsic harmony of the visual image, our intuitive ideas
concerning the regularity, symmetry and order prove here to be effective
control parameters of the process of visual determination of mathematical
regularities.
The problem next in complicacy is representability of numbers as
the sum of cubes. If we dealt not wich integers but with rational numbers,
mo« 2.3 the answer to the question would be the following.


n



~ .•
n

Theorem. (T. Raily 1825, Richmond 1930). Any rational number a is

.
representable as the sum of three cubes of rational numbers, namely:
t • •

• • .

t •
~

• ~
• ( a3 -3 6
) 3 ( -a3 +3 5a+3 6 )3 ( a2 +34a )3
a = 32a2 + 34a +36 + 32a2 +34a +36 + 32a2 +34a +36

• • • •
• •• ,
• , .
t We now proceed to the classical Waring problem. We fix r = 3 and
• • vary s, i.e. investigate the possibility of representing numbers n as the
t
~
I
n n;.
sum of s cubes: = nf + ... + We again display the information
• • • in the form of a carpet. In Fig.4.3.4 there are pictures (for s = 1,7,8)
J. (Gltl.lr) successively appearing with increasing s = 1,2,3,4,5,6,7,8,9. This
Fig. 4.3.3 In 0 2lr S material would generally be more convenient to illustrate by a cartoon
4.3 Visual Computer Geometry in the Number Theory 289

shot directly from the computer display. The first shots will be almost
all black, then the white field begins increasing. Finally, for s = 8 in
the white field there remain only two black squares which vanish for
s = 9.So for s = 9 the whole display is white! By the way, the last two
black squares in the eighth shot are numbers 23 and 239.

Computer hypothesis 2. Any natural number is representable as the


sum of nine cubes.

This is the well-known theorem by the German mathematician A. 1=1-


Wiferich which he proves in 1909.
r n

Computer hypothesis 3. At first (jar small s) there is very much black • •



.• •.
colour (and the set of black squares is obviously infinite), then there
becomes much less black colour (and the set of black squares becomes
(

obviously finite). And finally black colour vanishes altogether (the set of
black squares is obviously empty).


From the pictures obtained one can hypothesize that this situation (

•• •
may also hold for the other powers r = 4,5,6, .... The schematic ex-
pression of this hypothesis is presented in Fig. 4.3.5. Here s is the number
of summands in the sum (1), r 2: 2 is fixed exponent, N is the number L
of black (i.e. non-representable) natural numbers. Let r be fixed. Since 4= 1

the number of summands s increases monotonly (Fig. 4.3.5), there exists!+- r _ _ _u-n _----.
a certain value s = G(r) (of course, r-dependent) such that for s < G(r)
the black set is inifmite and for s 2: G(r) finite. In other words, G(r) is

the smallest number of summands in the sums (1) for which the black
set becomes finite for the first time. [
A similar boundary arises with a further increase of s, i.e. when the
number of elements of the black set goes over from finite to zero. This •
second boundary we denote by g(r). So, g(r) is the smallest number of
summands in the sum (1) for which the black set appears empty for the C
first time~
To say it differently, for s 2: G(r) it is only afinite amount of natural
numbers that is not decomposed into the sum nI + ... + n~, while for
s > g(r) all natural numbers are representable in the form nJ + ... +n~. 1= 8 Fig. 4.3.4
290 4. Visual Images in Some Other Fields of Geometry

Hilbert functions These two functions, G(r) and g(r), are called Hilbert functions. We
have presented our arguments, prompted by computer drawing, in favour
of the hypothesis that the number g(r) is finite. This does not, of course,
follow from its defmition.

Waring's conjecture Waring's hypothesis. For any power r ~ 2 there exists (and is finite)
a smallest number of summands g(r) such that for all s > g(r) the set
of natural numbers non-representable in the form nj +... +n~ is empty.
Here nj are non-negative integers.

Empirical hypothesis Waring formulated this hypothesis in 1770 on the basis of the La-
grange theorem and of the empirical assumptions of his predecessors
that any natural number can be represented as a sum of nine cubes and
nineteen fourth powers. There was no Wiferich theorem (1909) at that
time, but the corresponding empirical hypothesis on the sums of cubes
did exist.
o So, we have the equailities g(2) =4, g(3) = 9, g(4) = 19. The function
g(r) is seen to increase rapidly, and therefore it is a priori not excluded
that for a certain ro it will appear to be equal to infmity. This would
U 3· .. G('t) ... 1('T.) mean that for any arbitrarily large number of summands s there exists a
Fig. 4.3.5 sufficiently large natural n non-representable as the sum n = nro+...+n~o.
As a result of considerable attempts made by many mathematicians.
Hilbert-Waring theorem Waring's hypothesis was finally proved by D. Hilbert in 1909, and since
then it has been known as the Hilbert-Waring theorem.
The classical Waring's hypothesis was further on extended in differ-
ent directions. We shall not go into detail of these extensions, but only
mention the names of some mathematicians who obtained here some
further results: I.M. Vinogradov (1938, the Goldbach-Waring problem),
Hua Lo-gen (1937), Hua Lo-gen and R.E. Haston (1938), E.M. Right
(1934), B.I. Segal (1933), K.F. Rot (1951),V.I. Nechaev (1953), A.A.
Karatsuba (1962) and others.
There exists, however, one natural extension which we shall touch
upon here. In the course of computer experiments (A.A. Zenkin) on the
classical Waring problem, an interesting fact was unexpectedly discov-
ered. If in the sums (1) the condition "ni are non-negative integers" is
replaced by the condition "ni are natural (i.e. positive) integers", the
situation changes. It is of interest that this substitution of conditions was
4.3 Visual Computer Geometry in the Number Theory 291

primarily made by computer itself because it always understands too lit-


erally the desire of the experimenter to economize its memory. It began
discarding zero summands, i.e. the cases where ni = 0, which in its opin-
ion provided little information. It turned out that already in the simplest
case of the Lagrange theorem, i.e. for r = 2, instead of the Lagrange
theorem there appeared the cartoon shown in Fig.4.3.6. Ths distinction
is that as the number of summands s increases; the black field decreases
to the value s = 6, but with a further increase of s the number of black
squares remains unchanged! There remains a residue of seven numbers
of the form s+{1,2,4,5,7,10,13} for all s ~ 6. This visual empiri-
cal result (obtained due to an accidental misunderstanding between the
experimenter and the computer) has made the following contribution.

Fig. 4.3.6

Computer hypothesis 4. For r = 2 andfor all s ~ 6, any natural number


n ~ 1 is representable as the sums (1) with the restriction nj ~ I except
the numbers 1, 2, ... , s - 1 and numbers of the form s + z, where Z is
one of the numerals {1,2,4, 5, 7,10, 13}.

An additional and purposeful analysis of scientific literature dis-


covered the theorem of an American mathematician G. Poll which he
proved as far back as 1933 and which coincides litera111y with computer
hypothesis 4.
292 4. Visual Images in Some Other Fields of Geometry

The question naturally arose: What a strange set did the computer
distingush following Poll? It was decided to see, using computer, the
corresponding cartoon for cubes, i.e. for r = 3, which is n analogue of
the Wiferich theorem, but with the restriction on the summands of the
form nj ~ 1 in the sums (1). The result is demonstrated in Fig.4.3.7.

Computer hypothesis 5. For r = 3 and for all s ~ 14, any natural


number n is representable as the sum (1), where all nj ~ 1, except the
numbers for the following natural numbers: 1,2, ... , s - 1 and s +74
numbers of the form s + z, where Z are elements of the "strange" set
{1-6, 8-13, 15-20, 22-25, 27, 29-32, 34, 36-39, 41, 43-46, 50, 51, 53,
55, 57, 58, 60, 62, 64, 65, 67, 69, 71, 72, 74, 76, 79, 81, 83, 86, 88, 90,
93, 95, 97, 100, 102, 107, 109, 114, 116, 121, 123, 128, 135, 142, 149}.

As we see, there appears here a strange set which (and similar ones)
1. Lagarius called the exceptional set.
Computer hypothesis 5 was then given a rigorous mathematical proof
in Ref. [91]. This is a new mathematical result. We deal with an example
of a successful use of computer drawing for the formulation of a correct
mathematical hypothesis. Later on, many interesting properties of such
sets were investigated using computer, in particular, the hypothesis of
their finiteness was first formulated and then rigorously proved.

14
FiQ.4.3.7
Appendix 1. Visual Geometry of Some Natural
and Nonholonomic Systems

1.1 On Projection of Liouville Tori in Systems


with Separation of Variables

For simplicity we consider systems with two degrees of freedom, systems Natural
actually arising in nature, i.e. having Lagrangian of second degree in mechanical systems
velocities:
n n
L= L ajj(q)(iiQj - II(q) + L aj(q)qj
i,j=1 j=1

There are usually no linear summands (e.g. for the motion of a point
along a motionless surface) and then the system is called natural. There
is no conventional term to define the presence of linear terms. In some
cases we distinguish between reversible and i"eversible natural systems. Reversible
Reversibility implies that if q(t) is a solution, so is q(-t). The Hamilto- and irreversible
nian corresponding to our Lagrangian has the form natural systems

Consider completely integrable systems with Hamiltonian of particular


form for n = 2 (which means two degrees of freedom). We omit some
details referring the reader to the works by Tatarinov [93] and [94].
It is known that in the phase space of completely integrable systems Liouville foliation for
there appears fibre bundle (with singularities) of tori carrying condition- the integrable system
ally periodic windings, and if in the phase space their behaviour is more
or less the same, on the position manifold (configuration space) it is
294 AI. Visual Geometry of Some Natural and Nonholonomic Systems

~
>':":::';':":~':":""::"". rather varied because the phase tori and their windings can be projected

.: . . :.1)
. . .:.: . . .:. :.<.
· onto this manifold differently ([93], p.292). We mean here the projec-
·
'.

..... .. .... .'


.
· .::.:: ..... . tion (PI, P2, ql , q2) --+ (ql, q2). Let there be given a natural system with a
cyclic coordinate (ql = r, q2 = cp).

.. :.
.', ' ',' . : :, ", .. ::
" 1
H=-p2+~+II(r)
r., .
':.: ~.: .:.: .:/:: .:.:: :....: 2 r a(r)
.. , . '. " Then the torus is projected in a particularly simple way. One should put
..:............. :..
it onto the table and press it from above. The meridians cp = const go
Fig.A.1.1 over into segments orthogonal to the parallels r = const. The behaviour
of the trajectories of motion may vary as shown in Fig. ALl. The ap-
pearance of a linear term of the form Q(r )P<p in the Hamiltonian leads to
a peculiar twisting of the trajectories (Fig. A 1.2). In a sense, the winding
is projected here obliquely.
The situation is much more interesting in the so-called Liouville
Fig.A.1.2 systems:

For simplicity we consider a biharmonic oscillator

(2 2) +2I (1...2
1 PI +P2
H= 2 22)
wlql +W2q2

~! Then the integrals that are in involution are as follows:

~=~+~~=~, ~=~+~~=~
Fig. A.1.3 Along each coordinate there occurs harmonic oscillation with eigenfre-
quency WI, W2, such that the trajectory is a Lissajous figure in a rectangle
IqIi ::; .jCi/W"lq21::; .jCi./W2 (Fig.A1.3). Clearly, this rectangle is a
projection of the phase torus 1C1C2 = {~ =~, F2 = C2}, and the Lissajous
figures are projections of the windings. To imagine what it looks like, one
should glue a flat torus of a sheet of paper, plot on it a straight-line seg-
ment (a piece of winding) and glue the torus (Fig. A.1.4), see Ref. [93],
Fig. A.1.4 p.272. As we see, this gives a rectangle with a piece of billiard trajectory
inside it. Continuing this trajectory, we shall see something resembling
much a Lissajous figure. This is not a mere coincidence.
1.2 What Are Nonholonomic Constraints? 295

Theorem. A rectangle IICJ ,c2 can be homeomorphically mapped onto a


rectangle {I~i I ~ h}. so that the Lissajous figures will go over into
billiard trajectories making an angle of 7rI4 with the sides (Fig. A.l.5).
For the motion at a velocity I~il = 1 the quantities Tj = 27r IWi are
oscillation periods along the corresponding coordinate.

To prove this, it suffices to take ~i = llwi arcsinwiqi/..foi. This


theorem can be extended to arbitrary Liouville systems. But something
should be sacrificed. If we wish to see billiard trajectories, we have to
Fig.A.1.5
reject the simple meaning inherent in the size of the rectangle Ti. The
general theorem is formulated like this. Let there be given a Liouville Uouville system
system with the Hamiltonian

H = Pf/2+ Vi(ql)+~/2+ '\t2(q2)


fl(qd +h(q2)
and let there be fixed the constants of its first integrals H = h,

F =(b + hr J (h (pil 2+ Vi) - b (Pi12 +Vi) ) =c


(cf. [93], p.235, Exercise 52 and p.186 Theorem 2), so that

(Ii + h)2cfi12 - hi! + Vi = c, (Ii + idqj - hh +Vi = c


Suppose we are given a nonsingular phase torus feh = {H = h, F = c}
projected onto the rectangle

Then there exists a continuous map of this rectangle onto another one
of the form {O ~ 6 ~ T!(C, h), 0 ~ 6 ~ 1'2(c,h)}. such that the trajec-
tories of motion with given constants of the integrals are in one-to-one
correspondence with the billiard trajectories making an angle of 7r I4 Billiard trajectories
with the sides of the new rectangle. Nonh%nomic constraints

1.2 What Are Nonholonomic Constraints?


/~
~
Consider the following construction: along a horizontal plane there moves
a round continuous disc (Fig. A. 1.6). Its centre may occupy an arbitrary Fig.A.1.6
296 AI. Visual Geometry of Some Naturai and Nonholonomic Systems

position. Suppose the coordinates of the centre' are x, Y and the disc
may rotate arbitrarly (let the angle of rotation be <p), see Fig. A.1. 7.
Now fix a small skate, a blade of negligible length, from below to the
centre of the disc. Those who skate can imagine how easy it is to spin
Fig.A.1.7 and move forward or backward, whereas when pushed on the side you
will either fall or make a step because skates do not glide sideways.
1)' 1)'
Let us return to the disc and try to imagine how its centre can move.
, :::. . '. : :.' If we idealize the situation and use exact terms, we shall speak of the
direction of the velocity of the centre (Fig. A 1.8). The velocity of the
centre can be directed only along the blade. Let for defmiteness <p be
Ues .. ~ the angle between the blade and the x-axis. Then our requirement is that
.' :: . ~'. the velocity v= (x,iI) be always perpendicular to the normal to the blade
/!Q ::'.'

. ....
'
: . "
.
... :

"
n = (- sin <p, cos <p) or (v, n) = -x sin <p +iI cos <p = O. Restrictions of
this type imposed upon the velocities of parameter variation (in our case
Fig.A.1.8 there are the parameters x, y, <p) are called nonholonomic constraints.
Since the velocity of the centre is not arbitrary here, the question arises,
to what extent this lays restraint on our possibility to move the disc. More

~,.. ~"~"f,'
precisely, let there exist two positions of the disc where the parameters x,

.... --':..'1"'
1ft . . . ."!::. '
____~.
___ • __ ...... 2- y, <p acquire the values XI, YI ,<PI and X2, Y2, <P2 respectively (Fig. AI.9).
It is possible to transfer the disc from one position to another respecting
~ .' " , the nonholonomic constraint? The answer appears to be affirmative:
1 .:.'
, ,:
,
,
,I
:%'! :!'e.t
1) we put the blade alon the line of the centres,
2) glide from (XI,YI) to (X2,Y2),
Fig.A.1.9
3) turn the blade to obstain the angle <P2.
This type of nonholonomic constraint is called nonintegrable. We
do not discuss these concepts here and restrict ourselves to the example
given above. For the details see the papers by Tatarinov [93], [94].
An interesting nonholonomic constraint can be imposed on a body
with a motionless point (this is so-called Suslov's constraint in Wagner's
representation). First of all, how shall we realize this motionless point?
Imagine a spherical hinge: a mobile sphere glides along an immobile
Fig.A.1.10 one (Fig.Al.l0). The body can turn arbitrarily round the point 0 (for
the time being we forget about the support). The body is customarily
associated with a co-moving coordinate system 6,6,6, and the angular
velocity of the body w is decomposed in the vectors of the co-moving
frame: W = WI e~l + w2e~2 +w3e~3' If P is an arbitrary point of the body,
1.3 The Variety of Manifolds in the Suslov Problem 297

its velocity has the form Vp = [w, OP], and the coordinates of any point
are constant (in a co-moving coordinate system).
Now to the mobile sphere we fix a small blade of the same type as
in the preceding example. For definiteness we place it at the point P on It P.
the 6 -axis parallel to the 6-axis (Fig. A.1.11). The velocity of the point
P has now to be directed along the 6-axis since vp is always orthogonal
to 0 P, and in this case to the 6 -axis. It remains to require that Fig. A.1.11

i.e. W3 = O. This is just Suslov's constraint.

1.3 The Variety of Manifolds in the Suslov Problem Suslov's constraint

We shall deal with integral manifolds in phase space. Consider a body


moving in an axisymmetric potential field. The symmetry axis of the Co-moving
field will be denoted by X3 (the third axis in a fixed coordinate system) coordinate system
and let ')'1, ')'2, ')'3 (where ,),T + ')'i + ')'~ = 1) be components of the vector
eX3 in the co-moving coordinate system. Then the potential energy of the
body II = II (')'1, ')'2, ')'3), and the equations of motion without constraint
have the form m+ [w, m] = [av/a,)" ')'], l' + [w,,),] = O,')'T + ')'i + ')'~ = I
(the Euler-Poisson equations). Here m = Aw is th(e ~gu~r m~me)ntum, Euler-Poisson equations

and we assume the operator A to have the form 0


A2 0 ,
o A3 0
i.e. the axes of the co-moving coordinate system to be the principal axes Rigid body in an
for mass distribution in the body. We write the equations of motion axisymmetric potential field
(without constraint) in more detail: Principal axes
in a rigid body
dWI all all
AI-+(A3 - A2)W2W3 =')'2- -')'3-
dt a')'3 a')'2

dW2 all all


A2-+(AI- A3)W3WI =')'3- -')'1-
dt a')'l a')'3
dw3 all all
A3- +(A2 - A 1)WIW2 = ,),1- - ')'2-
dt a')'2 a')'l
298 Ai. Visual Geometry of Some Natural and Nonholonomic Systems

d'Yl
dt =W3 "12 - W2'Y3 ,

d; =W2'Y1 - WI'Y2 , 'Yf +'Yi +'Y~ = 1 .


Motion with constraint Now impose the constraint WJ = o. Then from the equation of motion
there disappears the part of summands where W3 is a co-factor, and the
third equation containing W3 is simply needless. We obtain

dwl all
AI- ="12- - "13-,
all ~ all
A2- ="13- - "11- ,
all
dt Ert3 0"12 dt 0"11 0"13

d'Yl d'Y2 d'Y3


dt = -W2'Y3 , dt = WI'Y3 , dt = W2'Y - WI'Y2 ,

i+'Yi+'Y~=l .
This "derivation" of the equations of motion with constraint is for-
mally not absolutely lawful. The correct derivation (which we omit here,
d'Alembert-Lagrange see Ref. [94]) is based on the d' Alembert-Lagrange principle by virtue
principle of which the action of an additional force moment orthogonal to possi-
ble angular velocities, i.e. exactly along the mobile 6-axis, is respon-
sible for the constraint. The system of equations obtained specifies the
flow on the manifold 8 2 ('Yf +'Yi +'Y~ = 1) X ]R2(p, q). Assume now that
II = III ("11) + Ih <'Y2). Then the flow obviously possesses two first in-
i
tegrals FI = (1/2) . Alwy + Il2 ("12), F2 = (1/2) . A2W + Ild'YI). The
question is: what can the topological structure of two-dimensional mani-
2-dimensional common folds Ic,C2 = {FI = C), h = C2} C 8 2 x]R2 be? They are always compact
level surfaces may have (the sum of the integrals written down gives the energy integral) and,
genus from zero to five as shown by Okuneva in [95], [96], may have genus from zero to five.
Integration in The equations of motion are integrated in quadratures, so that we have a
quadratures for the clear distinction from those integrable Hamiltonian systems in which the
equation of motion common level surfaces of the integrals-may have only genus 1 (tori). To
prove this assertion, we make the change of variable dr ='Y3dt in each
of the hemispheres "13 > 0, "13 < o. Then in the equations of motion we
Uouville-Hamiltonian can single out the subsystem
system
dwl 1 dIl2 d'Y2
a; - - A2 d'Y2 ' -=WI
dr
1.3 The Variety of Manifolds in the Suslov Problem 299

dw2 I dIll
-=+--
dr A2 d"Y1'
The integrals will of course be the same. This is the Liouville Hamiltonian
system:
" I dIll " I dIl2
'1'1 = - - - '1'2 = - - -
A2 d'Yl ' Al d'Y2

Therefore, all the arguments of Sect. ALl are valid here. In particular, Fig.A.1.12
in the phase space we obtain a torus which is projected onto a rectangle
in the variables '1'1, '1'2, and the trajectories are qualitatively Lissajous
figures in this rectangle. But working in the hemispheres '1'3 ~ 0 we
should remember that actually 'Yf + 'Yi ~ 1, and therefore we should
take the intersection of the rectangle with the circle (Fig. A.1.12). The
genuine trajectory (while the point is on the hemisphere) will be only the
portion of the "Lissajous figure" which lies inside the circle. By virtue Fig.A.1.13
of symmetry of the equations (namely, '1'3 --+ -'1'3 for t --+ -t), the
trajectory 'Y(t) is symmetric about the circle '1'3 = O. In the projection
onto the plane 'YJ. '1'2 this means that (')'1 (t), 'Y2(t») reaches 'Yf + 'Yi = 1,
is reflected and returns along the same trajectory.

.€D'.
Ascending to the phase space, over the hemisphere '1'3 > 0 or '1'3 < 0
(over a truncated rectangle) we shall see a torus which will also be
truncated. One can make sure that in the situation shown in Fig. A.1.12 Fig.A.1.14
(when one angle is cut out of the rectangle) a disc is cut of the torus
(Fig.ALl3). By virtue of the above-mentioned symmetry, I c1c2 consists
of two symmetric pieces which are glued together along the boundary ::':'. ':.:: "

(Fig. A.1.14). We obtain a pretzel! (And not a torus). It can readily be :~ '0:' '0 . :"
•• I' •. "

verified that in other versions of intersectij)n of a rectangle with a circle ......... .

. '
"

00
we obtain manifolds shown in Fig. AI.IS. This problem has been studied
2
in detail by Okuneva [95], [96] who gave an exhaustive description of the "
qualitative picture of the behaviour of trajectories on integral manifolds. ...) ...J

O· Q'
4,.,~.j
:"
....f
~

Fig.A.1.15
Appendix 2. Visual Hyperbolic Geometry

-= -
2.1 Discrete Groups and Their Fundamental Region

In this Appendix we consider the discrete groups of motion of Euclidean


spaces, spheres and Lobachevskian spaces generated by reflections. The
· .A.2. 1 groups of motions of a space are called discrete if the images of any point
FIg
form a discrete set under the action of all elements of the group. Discrete
groups are characterized by the fact that they have a fundamental region
i.e. a set whose images cover under the action of the group the entire
space and do not intersect except, perhaps, at the points of the boundary.
Examples of discrete groups may be the following:
1. The group oftranslations ofa Euclidean plane by integer-valued
vectors in a fixed Cartesian coordinate system. A fundamental region of
this group is the square 0 ~ x ~ 1, 0 ~ y ~ 1 (Fig.A.2.1). The hexagon
(Fig. A.2.2), the curvilinear figure (Fig. A.2.3) are also fundamental re-
Fig. A.2.2 gions of the same group.
2. The group of rotations of a circle through angles multiple by
21'i / n, where n is a fixed integer. In this case any arc of the circle, equal
to 21'i / n, is a fundamental region.
3. The group of motions of a ring, consisting of two elements:
identical and central symmetry motions. Any semicircle, or one of the
yin-yang figures (Fig. A.2A), is a fundamental region.
These examples show that a fundamental region of a discrete group
can be chosen in more than one way, and may be rather multiform. But
for the groups of motion of a plane (space) a fundamental region can
Fig. A.2.3 always be chosen in the form of a convex polygon (polyhedron) ([97],
2.2 Discrete Groups Generated by Reflections in the Plane 301

Chap.4, Sect. 1). A convex polygon (polyhedron) is understood as an


intersection of a family of half planes (half spaces) which in our case
is everywhere finite. The general definition allows the polygon (polyhe-
dron) to have an infinite number of sides (faces), but in this case some
natural restrictions should be imposed upon the family of half planes
(half spaces).

2.2 Discrete Groups Generated by Reflections in the Plane Fig.A.2.4

We shall understand reflection as axial symmetry; the axis of symme-


try will be called a mirror. The discrete groups of motions of a plane,
generated by reflection, are distinguished from among all other discrete
groups of motions for simplicity of their geometrical description. More
precisely, a fundamental region of such a group can be chosen in the
form of a convex polygon all of whose angles are integer parts of 7r
and vice versa, each such polygon is a fundamental region for a certain
discrete group generated by reflections.
We shall describe how, knowing the group generated by reflections,
one can find its fundamental polygon of the indicated form. To this
end, we consider the mirrows of all reflections belonging to the group
(Fig. A.2.5). They divide the plane into convex polygons which we shall Fig. A.2.S
call cells. We shall prove that any cell is a fundamental region of the
group and has angles equal to an integer part of 7r. To begin with we
shall show that the image of a mirror under reflection from another
mirror is also a mirror (the mirror is understood here only as the mirror
of reflection belonging to a given group). We use here the following
notation; for any straight line II, reflection relative to II is denoted by
RIl. Let II] and II2 be two mirrors. Then RRIJl (II2) = RIll' RIl2 . RIll
(Fig. A.2.6) obviously belong to the given group since RIll and RIl2
belong to this group. Thus, mirrors and, therefore, cells exchange places
under the action of the group. Any cell is carried to the neighbouring one
via reflection relative to their common side which is a mirror. By means
of a composition of such reflections, which is the element of the group,
any cell can be carried into any other one. More accurate arguments
([98], Chap. 5, Sect. 3, Subsect. 3) show that a cell can be carried into Fig. A.2.6
302 A2. Visual Hyperbolic Geometry

itself only by a unit element of the group. The latter two statements just
Fundamental region imply that a cell is a fundamental region.
We shall now show that the angles of the cell are integer parts of II.
Consider an arbitrary angle of a cell P, limited to straight lines IIo and
III. Consider the sequence of straight lines given by the recurrent formula
IIn+1 = Riin (lIn-I) (n ~ 1). According to what has been proved above,
Mi"ors all IIn are mirrors. If we assume the initial angle not to be an integer
part of II, then at some moment we obtain a mirror intersecting a cell
(Fig. A.2.7), which contradicts the definition of a cell. Therefore, the
angle under consideration is an integer part of 7r.
On the other hand, any polygon with angles equal to integer parts of
Group generated 7r can be shown [99] to be a fundamental region for the group generated
by reflections by reflections relative to the sides of this polygon. Thus, the classifi-
cation of reflection-generated discrete groups of motion of a plane is
reduced to the classification of polygons with angles of the form 7r / n.
Coxeter polygons Such polygons are called Coxeter polygons. All Coxeter polygons can be
directly enumerated. All of them are presented in Fig.A.2.8. To solve an
analogous problem in a three-dimensional space and in spaces of higher
dimensions, it is useful to introduce some new concepts.

Fig.A.2.7

1M1OP1L~~
(It-) (5) (6)

(1 ) (2 ) (3)
~ (1)
~.(8 ) (9 )
i.iai pfane
Fig.A.2.8
2.3 The Gram Matrix and the Coxeter Scheme 303

2.3 The Gram Matrix and the Coxeter Scheme

Coxeter polygons can be defined by the Gram matrix of a set of normal-


ized vectors orthogonal to the sides and directed outwards. This matrix
will be called the Gram matrix of a polygon. Gram matrix
Another way to define Coxeter polygons is the Coxeter scheme. The of a polygon
Coxeter scheme is a graph with vertices corresponding to the sides of Coxeter scheme
the polyhon. Two vertices are not joined if the corresponding sides are
perpendicular. If the sides intersect at an angle 7r In, the vertices are
joined by an edge of multiplicity n - 2 (or by a simple edge labelled
n). If the sides are parallel, the vertices of the scheme are joined by a
bold-faced edge (or by an edge labelled (0).
Obviously, a Coxeter polyhedron is uniquely defined, with an accu- Coxeter polyhedron
racy to the motions of the plane, by the Gram matrix or by the Coxeter
scheme.
The Gram matrices of the polygons depicted in Fig. A.2.8 are pre-
sented in Table A.2.2 gives their Coxeter schemes. Naturally, to the case
of Fig. A.2.8(9), i.e. to a polygon having not a single side there corre-
sponds a zero-dimensional Gram matrix and an empty Coxeter scheme.

Table A.2.1

o
(1)
~
(2)
()IIIIIIO
(3 )
0
L(6 )
o " () ~
(8)

Table A.2.2

(1) (:1-:
o
~~
0 i -1
o 0 -1 1
l
(I )
(j )

2.4 Reflection-Generated Discrete Groups in Space Reflection-generated


discrete groups
Reflection in space will be understood as symmetry about a plane. Simi- Reflection in space
larly to the case ofthe groups on a plane, the reflection-generated discrete
304 A2. Visual Hyperbolic Geometry

groups of motions of a space are defmed by their fundamental region


which may be any convex polyhedron with dihedral angles of the form
Coxeter polyhedra 7r/n. Such polyhedra are called Coxeter polyhedra. The concepts of the
Coxeter scheme and the Gram matrix are extended, without any essen-
tial changes, to Coxeter polyhedra of any dimension. One-dimensional
Coxeter polyhedra are, naturally, a segment, a ray, a straight line. Their
Coxeter schemes will be 0- - - -0, 0 and an empty scheme, respec-
tively.
Direct product of A direct product of Coxeter polyhedra is, obviously, a Coxeter poly-
Coxeter polyhedra hedron. Inversely, if a Coxeter polyhedron is represented as the direct
product of two polyhedra of lower dimension, the latter are also Coxeter
Decomposable polyhedra. A polyhedron representable in the form of the direct product
Coxeter polyhedra of polyhedra of lower dimensions will be called decomposable. Among
the Coxeter polyhedra of Fig. A.2.8, decomposable are 8(1), 8(2), 8(3),
8(4) for n = 2; 8(5) and 8(9).
Classification of The definitions of the Coxeter scheme, the Gram matrix and the di-
Coxeter polyhedra rect product of polyhedra imply equivalence of the following conditions:
1) The Coxeter polyhedron falls into the direct product of Coxeter
polyhedra of lower dimensions.
2) The set of vectors orthogonal to the faces of a polyhedron falls
Orthogonal subsets into a union of pairwise orthogonal subsets.
3) The Gram matrix ofa polyhedron has a cell-diagonal form under
a permutation of rows and the same permutation of columns.
Subschemes in a 4) The Coxeter scheme of a polyhedron falls into a union of pair-
Coxeter scheme wise disconnected subschemes.
The classification of Coxeter polyhedra reduces, obviously, to the
classification of undecomposalbe Coxeter polyhedra.

Exercise 1. Suppose in an n-dimensional Euclidean space we are given


a set of vectors forming pairwise non-acute angles, which does not fall
into a union of pairwise orthogonal subsets. Then the vectors of the set
may have no more than one linear deprendence. In particular, the set
contains no more than n + 1 vectors. All the coefficients of the linear
dependence, if it does exist, are positive.
2.4 Reflection-Generated Discrete Groups in Space 305

Since the angles between the vectors orthogonal to the faces of a


Coxeter polyhedron and directed outwards are equal to 7r-7r In, They are
not acute, and the statement of Exercise 1 implies that undecomposable Undecomposable
Coxeter polyhedra are simplexes or simplicial cones. To the simplexes (irreducible)
there correspond non-negative definite degenerate Gram matrices whose Coxeter polyhedron
rank is smaller than the matrix dimension by unity. To the simplicial
cones there correspond positive definite Gram matrices. Note also that
all the diagonal elements of the Gram matrix of a Coxeter polyhedron are
equal to 1, while the non-diagonal elements are either equal to -lor have
the form - cos 7rI n. All the matrices satisfying the above-mentioned
conditions are classified [99], and thus there exists a complete description Complete classification
of Coxeter polyhedra in Euclidean spaces of any dimension. The schemes of Coxeter polyhedra
of undecomposable Coxeter polyhedra in Euclidean spaces of dimension in Euclidean space ]Rn
not higher than three are presented in Table A.2.3.
Table A.2.3
Si mpl'ic"q'[ CMfS in E It
n Simpi4r.fS ill Ell (si"'pR*~ ill $")

.{ ~ 0

2 \lo=c:::oCHIiD /'I.
0--0

3 n >=' 0-0-0
o-a:;::D
()-(E3)

Reflection-generated discrete groups with positive definite Gram ma- Positive definite
trices have the direct product of simplicial cones as their fundamental Gram matrices
region. These groups have a motionless (fixed) point and can be in- Simplicial cones
terpreted as reflection-generated discrete groups of motions of a sphere
(with centre at the fixed point). The Coxeter schemes of these groups are
given in the right-hand column in Table A.2.3.
306 A2. Visual Hyperbolic Geometry

Exercise 2. Prove that the discrete group of motions of a Euclidean space


has a (common) ftxed point if and only if it (this group) is ftnite.

Figure A.2.9 shows all Coxeter polyhedra on a two-dimensional


sphere, their Coxeter schemes and sphere subdivision into cells under
the corresponding discrete group.

Exercise 3. Count the number of cells in each case.

o--!!:--o 0

Fig.A.2.9

Lobachevskian or 2.5 A Model of the Lobachevskian Plane


hyperbolic plane
We shall consider a three-dimensional pseudo-Euclidean space V 2,1 with
the scalar product defined by the quadratic form

(x, x) = -x~ + xI + x~ .

The set c = {x E V2,1 : (x,x) < 0 is an open cone which consists of two
connected components C+ and C- (Fig. A.2.1 0). A set of rays emanating
Model from zero and lying in c+ is a model of the Lobachevskian plane A2.
2.5 A Model of the Lobachevskian Plane 307

Fig.A.2.10

The rays lying on the boundary of C+ are called infinitely remote points Infinitely
of the Lobachevskian plane. remote points
Straight lines in this model consists of rays lying in the intersection
of a certain two-dimensional subspace of the space V 2,1 with c+ if this
subspace is non-empty. Each straight line in A2 can be defined by the
vector of the space V 2,1 orthogonal to this subspace. Given this, the
condition that the intersection of this subspace with C+ is non-empty is
equivalent to the fact that the restriction of the quadratic form to it has a
signature (1,1), which is in turn equivalent to the fact that the orthogonal
complement to it is spanned by a vector with a positive scalar square.
Thus, each vector x E V 2,1, such that (x, x) > 0 defines a straight line
IIx C A2.
It is convenient to draw the cross-section of the cone C+ which
intersects each ray lying in the cone at exactly one point. Such a cross-
section can also be interpreted as a model of the Lobachevskian space.
An example of such a cross-section is that by the plane Xo = 1. In this
Fig.A.2.11
case, the unit circle is a model of the Lobachevskian plane and the chords
of this circle are the straight lines. This model is called the Klein model. Klein model of a
It shows very well three possible cases of mutual disposition of two hyperbolic plane
straight lines on a Lobachevskian plane. They can intersect each other Intersection "at infinity"
(Fig.A.2.l1(1», be parallel (Fig. A. 2.1 1(2» or diverge (Fig.A.2.11(3». or "behind infinity"
In these cases, the lines of intersection of two-dimensional planes of the
space V2,1 specifying these straight lines, lie in C+, lie on the boundary
of the space and do not lie in its closure, respectively. We shall also say
that two straight lines on a Lobachevskian plane intersect at a proper
point, "at infinity" or "behind infinity" depending on which of the three
cases is realized.
308 A2. Visual Hyperbolic Geometry

Motions of a The motions of the Lobachevskian plane in the model described


hyperbolic plane are tranformations corresponding to the c+ -preserving pseudoorthogonal
transformations of the space V 2,1.
The reflections (relative to straight lines) are motions correspond-
Pseudo-orthogonal ing to pseudoorthogonal reflections relative to the two-dimensional c+-
transformations intersecting subspaces of the space V2, 1• Such a reflection Re relative to
the two-dimensional subspace orthogonal to the vector e is given by the
formula
Rex = x _ 2(e, x) e .
(e, e)

The invariants of relative position of straight lines on a Lobachevskian


plane are the angle between them in case they intersect and the distance
between them in case they diverge. The case of parallel straight lines is
limiting 'for both these cases. The angle between parallel lines and the
distance between them are assumed to be equal to zero.
On the other hand, from the description of motions in the model
in question it follows that such an invariant is the scalar product of the
vectors of the space V 2,1, which define straight lines (these vectors are
assumed to be normalized and agreeably oriented).
The relation between these invariants is given as follows:

Fig.A.2.12
Note that the first formula coincides with that for the case of Euclidean
plane. The minus sign appears because (eJ,e2) = 7r - (IIe;,rre
2 )'

Convex polygons on 2.6 Convex Polygons on the Lobachevskian Plane


the hyperbolic plane
The definition of a convex polygon on the Lobachevskian plane does not
differ from the case of Euclidean plane. Each convex polygon in A2 can
be given by the set of vectors in V2,1 specifying the sides of the polygon.
In order that this set of vectors can be uniquely restored from the polygon,
we assume the vectors to be normalized and directed outward the cone
2.7 Coxeter Polygons on the Lobachevskian Plane 309

constructed in V 2,1 over the given polygon (Fig. A.2.12). The polygon is
defined, up to the motions, by the Gram matrix of this set of vectors.
The sum of the angles of a polygon on the Lobachevskian plane is
less than the sum of the angles of a Euclidean polygon with the same
number of sides. The difference between these sums of angles is equal
to the area of the polygon on the Lobachevskian plane [97]. Therefore,
as distinguished from the Euclidean plane, the area of a polygon on the
Lobachevskian plane is fmite if some of its adjoint sides are parallel, in Polygons of a
other words, if some of the vertices lie at infinity (Fig.A.2.13). However, finite area and of
if at least one vertex lies behind infmity, the area of the polygon is an infinite area
infinite. Indeed, in the latter case one can construct a straight line lying
entirely in the polygon so that the polygon contains the whole of the
half plane restricted by this straight line (Fig. A.2.l4). Infinity of the
area of the polygon now follows from infinity of the area of the whole
Lobachevskian plane (the latter follows, for example, from the existence
of an infinite diescrete group of motions on the Lobachevskian plane).
Thus, a convex polygon has a finite area on the Lobachevskian plane if
and only if it is a convex hull of a finite number of points which are
either proper or lie at infmity.

2.7 Coxeter Polygons on the Lobachevskian Plane

As in the Euclidean plane, discrete reflection-generated groups of mo- Discrete


tions of the Lobachevskian plane are in one-to-one correspondence with reflection-generated
Coxeter polygons. Coxeter polygons on the Lobachevskian plane can groups of motions of
also be defined by the Gram matrix or the Coxeter scheme. a hyperbolic plane
The Coxeter scheme is somewhat modified as compared with the Eu-
clidean case. Namely, we shall join by a bold-faced edge only vertices
corresponding to parallel sides. If the sides diverge, the corresponding
vertices will be joined by a dashed edge. In case it is necessary to recon-
struct the polygon from the Coxeter scheme, we shall label the dashed
edge the same as the corresponding element of the Gram matrix.

Exercise 4. Prove that if the two straight lines, on which some sides of a
Coxeter polygon lie, intersect then their intersection is a vertex. In other
words, continuations of non-adjoint sides do not intersect.
310 A2. Visual Hyperbolic Geometry

Instruction. Make use of the fact that all the angles of a Coxeter
polygon are non-obtuse and that the sum of the angles of an arbitrary
triangle on the Lobachevskian plane is less than 7r.

D
It is now obvious that to a pair of adjoint sides of a Coxeter poly-
gon there corresponds a subscheme of its Coxeter scheme of the form
o 0 or 0 n 0 (where n ~ 3), to a pair of sides in-
tersection at an infinitely remote vertex there corresponds a subscheme
*3 o 0 and to a pair of non-adjoint sides there corresponds a subscheme
2~~., ~., ~J LOO 0----0·
This implies that restricted Coxeter triangles are specified by schemes
i+!..f..J:{i
.fc, ~, 'i<J of the form shown in Fig.A.2.i5a; the schemes for rectangles are shown
in Fig. A.2.l5b, for pentagons in Fig. A.2.15c, ect. Admitting that k - i
Finite area can assume the value 00, we obtain the schemes of finite-area Coxeter
Coxeter polygons polygons. Admitting dashed edges instead of ordinary or multiple ones,
we obtain the schemes or arbitrary Coxeter polygons.
It should be noted that for a polygon to be specified by its own
, \ /
.- Coxeter scheme, it is necessary that the dashed edges be marked. This
t. ?, -k1. can be done not in an arbitrary way. It is necessary that the Gram matrix
/ " of the polygon have rank three. This condition implies that a finite-area
f." Coxeter n-gon on a Lobachevskian plane is specified by its angles (which
2{.~, ~.,-l-" ¥oO must satisfy some inequalities) up to n-3 continuous parameters.
~i t-2. simuftanllouslj
2.8 Coxeter Polyhedra in the Lobachevskian Space

The model of the Lobachevskian plane constructed in Sect. A.2.5 admits


of natural extension to the case of Lobachevskian space. The initial space
in which the model of the three-dimensional Lobachevskian space A3 is
S' constructed is a pseudo-Euclidean space V 3,1 with the scalar product
u~,I.,k./l;.,I<t(.Oa given by the quadratic form
Fig.A.2.15 (x, x) = -x~ +xI +x~ +x~
Minkovski space The space V3,1 is also referred to a Minkowski space. The space A3 is
realized as a set of rays emanating from zero and lying in one of the two
connected components c+ of the cone {x E V3,1 : (x, x) < o}. As in
Klein model of a the case of the plane, from this model one can obtain the Klein model
hyperbolic 3-space of Lobachevskian space. To this end one should draw a cross-section of
2.8 Coxeter Polyhedra in the Lobachevskian Space 311

the cone by a hyperplane Xo = 1. The Lobachevskian space in the Klein


model is a Euclidean ball and the planes in this space are cross-sections
of the ball by Euclidean planes.
The definitions of motions, reflections, Coxeter polyhedra, their
Gram matrices and Coxeter schemes are obtained using natural modi- Gram matrix and
fications of the corresponding two-dimensional concepts. Coxeter scheme
For each edge (vertex) of a fixed Coxeter polyhedron we consider
a Coxeter polyhedron limited to all the faces of the initial polyhedron
which pass through this edge (vertex). The Gram matrix and the Coxeter
scheme of the polyhedron obtained will be called the Gram matrix and
the Coxeter scheme of the edge (vertex).
Each vector e E V 3,1, such that (e, e) = 1, specifies the plane
II2 C A3 as the set of rays lying in the intersection of the orthogo-
nal complement of e in V3,1 with C+. Inversely, each such vector is
determined by the plane from A3 uniquely up to a multiplication by -I.
Two planes IIe! and IIe2 intersect in A3 along a straight line if
and only if I(el, e2)1 < I. In this case cos (IIe;-IIe 2) = -(e), e2). The
planes IIe! and IIe2 intersect at an infinitely remote point provided that
l(eJ, e2)1 = I. And fmally IIe! and IIe2 diverge provided that l(eJ, e2)1 >
I. In this case ch {I (IIep IIe2 ) = -(eJ, e2). All this can be readily deduced
from the corresponding formulas for A2. Figure A.2.16 presents different (.3 )
cases of the relative position of two planes in the Klein model. Note that Fig. A.2.16
these cases correspond to a positive definite, a degenerate non-negative
definite and an indefinite Gram matrices of the set of vectors {e), e2}.
Similarly, from the signature of the Gram matrix one can find Signature of the
whether several planes intersect at one point. More precisely, IIe!, IIe2 , Gram matrix
... ,IIek , k ~ 3 intersect at one point from A3 ifand only ifGr(el, ... ,ek)
is a non-negative definite matrix of rank 3. Indeed, IIe! n.. .nIIek is an or-
thogonal complement in V 3,1 of the subspace spanned by el, ... ,ek' For
this intersection to determine a point in A3, it is necessary and sufficient
that it should be one-dimensional positive definite since the signature of
the space V3,1 is equal to (3, 1). One can prove in a similar manner that
IIe!, ... IIek intersect at an infinitely remote point of the space A3 if and
only if Gr(e) , ... , ek) is a non-negative definite matrix of rank 2.
We shall apply the results obtained to the planes of faces of Cox-
eter polyhedra. Since (see Sect. A.2.4) positive defmite Gram matrices
312 A2. Visual Hyperbolic Geometry

correspond to spherical Coxeter polyhedra and non-negative defInite to


Spherical Euclidean ones, it follows that the Gram matrix of a proper vertex of
Coxeter polyhedra a Coxeter polyhedron in the Lobachevskian space is the Gram matrix
of a certain spherical Coxeter polygon and the Gram matrix of an in-
Euclidean fInitely remote vertex is that of a Euclidean Coxeter polygon. Therefore,
Coxeter polygon the Coxeter scheme of any proper vertex of a Coxeter polyhedron in the
Lobachevskian space in a union of connected components; they are listed
in the right-hand column of Table A.2.3 and the data for an infmitely
Proper vertices remote vertex are listed in the left-hand column. In particular, all proper
vertices of the Coxeter polyhedron have a simplicial form, i.e. at such
Infinitely remote a vertex there intersect three faces, while at an infinitely remote vertex
vertex three or four faces may intersect.
The inverse is valid too. To each subscheme of the scheme of a
Coxeter polyhedron, which falls into connected components from the
right-hand column of Table A.2.3 there corresponds a proper vertex,
while to a subscheme which falls into connected components from the
left-hand column of Table A.2.3 there corresponds an infInitely remote
vertex. To each subscheme of the form 0 n 0 (2 ::; n < (0) there
corresponds an edge.
This follows from the result obtained by Andreev [100] and pre-
sented below. This result is an extension of Exercise 4 to three dimen-
Acute-angled sions. We call a polyhedron acute-angled if all of its bihedral angles are
polyhedron non-obtuse (in particular, all Coxeter polyhedra are acute-angled). In an
acute-angled polyhedron in A3:
a) the planes of disjoint faces do not intersect;
b) if the planes of several faces intersect at a proper or at an in-
fInitely remote point of the space A3, this point is a vertex of the poly-
hedron.
Coxeter scheme Speaking of the Coxeter scheme signature we shall mean the signa-
signature ture of the corresponding Gram matrix. In the same sense we shall speak
Positive definite, of a positive defInite, a non-negative defInite or an indefInite Coxeter
non-negative definite, scheme.
indefinite The simplest Coxeter polyhedra in A3, i.e. the limited simplexes
Coxeter schemes (tetrahedra) can be easily enumerated. Their Coxeter schemes should
satisfy the following conditions:
a) They should not be non-negative defInite.
2.8 Coxeter Polyhedra in the Lobachevskian Space 313

b) Any of their proper subschemes is positive defmite. o_---a:===:a:!1==::0"


c) Their rank: is equal to 4. (1 )

D C
All such schemes are presented in Table A.2.4.
TableA.2.4 (2) (J)

'&3 FA'
C ~(t,.)

Fig.A.2.17

The schemes of Fig. A.2.17 show unrestricted finite-volume tetrahe-


dra (as in the planar case, the finiteness of the volume of a polyhedron
is equivalent to the fact that all of its vertices are either proper or lie at
o ~-~
CO (2)

~--<> D>
infmity). For example, the tetrehedron A.2.l7(1) contains a vertex at in-
finity, which is specified by the faces -2, 3, 4, the rest of the vertices are
proper. The tetrahedron A.2.l7(2) contains two infinitely remote vertices, (3) (4)
and all the vertices of the tetrahedron A.2.l7(4) are infmitely remote.

Exercise 5. Find all such tetrahedra. ~4


(S)
[}
(6)
The scheme ofFig.A.2.l8(J) is an infinite-volume tetrahedron since
the subschema is indefinite, and therefore the corresponding Fig.A.2.18
vertex lies behind infinity. But if we add another face to the tetrahedron
(which will cut the vertex), we shall obtain the scheme of a restricted Restricted
trigonal prism (Fig. A.2.l8(2)). In Fig. A.2.l8(3}-(6) one can see other trigonal prisms
examples of restricted trigonal prisms. The vertices joined by a dashed
edge correspond to the bases of the prisms.
Makarov [101 ]-[ 104] has proposed some geometric constructions
which help in building many examples of Coxeter polyhedra in
Lobachevskian space.
Note that for all the schemes of Fig.A.2.l8 the marks on dashed
edges are uniquely determined by the condition imposed on the rank of
the Gram matrix. The same is valid for Coxeter polyhedra of another Dihedral angles
combinatorial structure. E.M. Andreev has proved that an acute-angled of acute-angled
finite-volume polyhedron in Lobachevskian space is uniquely defined finite-volume polyhedron
314 A2. Visual Hyperbolic Geometry

by its dihedral angles. For a polyhedron to exist, its angles should sat-
isfy some natural restrictions in the form of inequalities [105], [106].
Andreev's result can be interpreted as the classification of acute-angled
polyhedra (and, in particular, Coxeter polyhedra) of fmite volume in
Lobachevskian space. This classification does not, however, provide an
explicit representation of these polyhedra (an explicit representation is
understood here as the defmition of a polyhedron by a set of vectors in
V 3,J). Indeed, to find all the elements of the Gram matrix of a polyhe-
dron from its bihedral angles, it suffices to solve a system of algebraic
equations. One of the ways to find an explicit representation of some
Coxeter polyhedra is discussed in the subsection to follow.

Discrete groups 2.9 Discrete Groups of Motions of Lobachevskian Space and


of motions of Groups of Integer-Valued Automorphisms of Hyperbolic
hyperbolic space Quadratic Forms
Integer-valued
automorphisms We consider an arbitrary integer-valued quadratic form of the signa-
of hyperbolic
ture (3,1). Over the field of real numbers this form is equivalent to the
quadratic forms
form -xij + XI + x~ + x~, and therefore the group of its automorphisms
can be interpreted as the group of motions of a Lobachevskian space.
The subgroup of integer-valued automorphisms is discrete, it contains
a maximal reflection-generated subgroup. Vinberg [107] has found the
Fundamental polyhedron algorithm for constructing the fundamental polyhedron of the maximal
reflection-generated subgroup in an arbitrary discrete group of motions.
This algorithm can be applied to groups of integer-valued automorphisms
of hyperbolic quadratic forms. Shaikheev [108] has found a series of
Coxeter polyhedra of a rather complicated combinatorial structure. M.K.
Bianchi groups Shaikheev has considered extended Bianchi groups which can be defined
Integer-valued as the groups of integer-valued automorphisms of quadratic forms
quadratic forms
-2XJX2 +2x~ +2mx~, for m == 1,2(mod4) and
-2XJX2 +2xJ +2X3X4 + m;J x~, for m == 3(mod4) ,
where m is a positive integer free of squares. He has shown that for
m ~ 30 the fundamental polyhedron of this group has a finite volume
if and only if m f 22, 23, 26, 29. This fundamental polyhedron has been
explicity found in all the cases.
2.10 Reflection-Generated Discrete Groups 315

We shall give another example reported by E.B. Vinberg. It comes


out of restriction of the quadratic form

to the hyperplane Xo + Xl + X2 + X3 = O. The fundamental polyhedron of


the maximal reflection-generated subgroup in the group of integer-valued Maximal
automorphisms of this form is combinatorially organized as a simplex reflection-generated
with two cut-off vertices (Fig. A.2.19; the figure also presents its Coxeter subgroup
scheme). The vectors defining this polyhedron are as follows:
~, I
v) = -e) +2e2 - e3; V2 = -e2 +e3 ; '0--<1
S' 3
V3 = -e4 ; V4 = eo +4e) - 2e2 - 2e3 ;
Vs = eo + 3e4 ; V6 = eo + 2e) - 2e3 + 2e4 ,
Fig.A.2.19
where eo, e), e2, e3, e4 is the basis in which the form is as indicated above.

2.10 Reflection-Generated Discrete Groups in Multidimensional


High-Dimensional Lobachevskian Spaces hyperbolic spaces

The model of an n-dimensional Lobachevskian space is constructed as in


the two- and three-dimensional cases in an (n + I)-dimensional pseudo-
Euclidean space with a scalar product given by the quadratic form

Reflection-generated discrete groups are in one-to-one corrrespondence


with Coxeter polyhedra which are defined by their Gram matrices or Cox-
eter schemes. The combinatorial structure of an n-dimensional Coxeter Combinatorial structure
polyhedron is established from its Coxeter scheme as follows. To proper of an n-dimensional
vertices of a polyhedron there correspond positive definite subschemes Coxeter polyhedron
of rank n, to edges those of rank n - 1, to two-dimensional faces those
of rank r - 2, etc. To infinitely remote vertices of the polyhedron there
correspond degenerate definite non-negative subschemes of rank n.
Owing to the existence of the complete classification of positive
definite and degenerate non-negative definite schemes [99], one can find
Coxeter polyhedra of a given combinatorial structure.
316 A2. Visual Hyperbolic Geometry

All restricted Coxeter simplexes have been enumerated by Lanner


[109]. They exist only in a space of dimension not higher .than four.
Finite-volume Coxeter simplexes (i.e. admitting vertices at infinity) exist
Simplicial prisms in spaces of dimension less than or equal to nine. Restricted simplicial
prisms exist in spaces of dimension lower than five. All of them are
decribed by Kaplinskaya [110].
As distinct from the two- and three-dimensional cases, Coxeter poly-
hedra in Lobachevskian spaces of higher dimensions exist when bihedral
angles are not arbitrarily given. These angles are interconnected by some
relations. The question of classification of finite-volume Coxeter poly-
hedra in An for n ;::: 4 remains open. It is known that for large n such
polyhedra do not exist. The following theorems are valid.

Theorem 1. (Vinberg [Ill]). In An for n ;::: 30 there exist no restricted


Coxeter polyhedra.

Theorem 2. (Prokhorov [112]). In An for n ;::: 996 there exist no finite-


volume Coxeter polyhedra.

Both these estimates can obviously be improved. Examples of re-


stricted Coxeter polyhedra are known only up to seven dimensions [113]
and examples of finite-volume polyhedra up to dimension 19 [114]. These
record figures have been obtained in consideration of groups of integer-
valued automorphisms of hyperbolic quadratic forms. Integer-valuedness
can be understood here in the sense of a ring of integers of a certain field
of algebraic numbers. In case this field does not coincide with the field
of rational numbers, then for the group of integer-valued automorphisms
of a quadratic form to be discrete, some additional conditions should
necessarily be imposed upon this form [115].
Examples of Coxeter polyhedra in Lobachevskian spaces are also
considered in Refs. [107], [116], [115], and [117]. The latter is a survey
of the subject in question and gives an extended list of references.
References

[1] Hilbert, D., Cohn-Vossen, S.: Anschauliche Geometrie. Springer-Verlag:


Berlin Heidelberg New York 1932
[2] Smale, S.: Topology and Mechanics. Invent. Math. 10 (1970)
[3] Poston, T., Stewart, I.: Catastrophe Theory and its Applications. Pitman
Books Inc.: London, 1978
[4] Hildebrandt, S., Tromba, A.: Mathematics and Optimal Form. Scientific
American Library. Scientific American Books, Inc. New York, 1985
[5] Novikov, S.P.: Hamiltonian formalism and a multi-valued analogue of the
morse theory. Usp. Mat. Nauk 37, No.5 (1982) 3-49 (Russian)
[6] Novikov, S.P.: Variational methods and periodic solutions or Kirchhoff
type equations. II. Funkts. Ana!. Prilozh. 15, No.4 (1982) 3-49 (Russian)
[7] Arnold, V.I.: Mathematical Methods of Classical Mechanics. Nauka:
Moscow, 1974. English trans!.: Grad. Texts Math. 60. Springer-Verlag:
New York Berlin Heidelberg, 1978
[8] Nitsche, J.: Vorlesungen fiber Minimalfliichen. Springer-Verlag: New
York Berlin Heidelberg, 1975
[9] Banchhoff, Th., Strauss, Ch.: Complex Function Graphs, Dupin Cyclides,
Gauss Map, and Veronese Surface. Computer Geometry Films. Brown
University: Providence, 1977.
[10] Mandelbrot, B.B.: The Fractal Geometry of Nature. W.H. Freeman and
Company: New York, 1977
[11] Francis, G.K.: A Topological Picturebook. Springer-Verlag: New York
Berlin Heidelberg, 1987
[12] Penrose, L.S., Penrose, R.: Impossible objects: a special type of illusion.
Brit. J. Psycho!. 31, 31-33
[13] Peitgen, H.-O., Richter, P.H.: The Beauty of Fractals. Images of Complex
Dynamical Systems. Springer-Verlag: Berlin Heidelberg New York, 1985
[14] Kolmogorov, A.N.: The Main Concepts of Probability Theory. 2nd edi-
tion. Nauka: Moscow, 1974 (Russian)
[15] Gnedenko, B.V.: The Course in the Theory of Probability. 4th edition.
Nauka: Moscow, 1965 (Russian)
318 References

[16] Shiryaev, A.N.: Probability. Nauka: Moscow, 1979. English trans!.: Grad.
Texts Math. 95. Springer-Verlag: New York Berlin Heidelberg, 1984
[17] Chentsov, N.N.: Statistical Rules and Optimal Conclusions. Nauka:
Moscow, 1972 (Russian)
[18] Maslov, V.P.: Perturbation Theory and Asymptotical Methods. Moscow
State University Press: Moscow, 1965 (Russian)
[19] Zakharov, V.E., Faddeev, L.D.: Korteweg-de Vries equation as a com-
pletely integrable Hamiltonian system. Funkts. Anal. Prilozh. 5, No. 4
(1971) 18-27 (Russian)
[20] Gelfand, I.M., Cherednik, I.V.: Abstract Hamiltonian formalism for clas-
sical Yang-Baxter equations. Usp. Mat. Nauk 38, No.3 (l979) 3-21
(Russian)
[21] Kozlov, V.V.: The Methods of Qualitative Analysis in Rigid-Body Dy-
namics. Moscow State University Press: Moscow, 1980 (Russian)
[22] Manin, Yu.I.: A Course of Mathematical Logic. English trans!.: Grad.
Texts Math. 53. Springer-Verlag: New York Berlin Heidelberg, 1977
[23] Manin, Yu.I.: Mathematics and Physics. Znanie: Moscow, 1979 (Russian)
[24] Fomenko, A.T., Fuchs, D.B., Gutenmacher, V.L.: Homotopic Topology.
Moscow State University Press: Moscow, 1969. English trans!.: Akademia
Kiad6: Budapest, 1986
[25] Fomenko, A.T., Fuchs, D.B.: A Course in Homotopic Topology. Nauka:
Moscow, 1989 (Russian)
[26] Fomenko, A.T.: Mathematical Impressions. Amer. Math. Society: Provi-
dence, R.I., 1990
[27] Dubrovin, B.A., Fomenko, A.T., and Novikov, S.P.: Modern Geometry.
Methods and Applications. Nauka: Moscow. 2 Parts: 1979, 1984. English
trans!. of Part I: Springer-Verlag: New York Berlin Heidelberg, Grad.
Texts Math. 93, 1984, and Grad. Texts Math. 104 (1985). Part II: Grad.
Texts Math. 124, 1990
[28] Fomenko, A.T.: Differential Geometry and Topology. Moscow State Uni-
versity Press: Moscow, 1983. English trans!.: Plenum Publishing Corpo-
ration: New York London, 1987
[29] Fomenko, A.T.: Variational Problems in Topology. Moscow State Uni-
versity Press: Moscow, 1984. English trans!.: The Geometry of Length,
Area and Volume. Gordon and Breach: London, 1990
[30] Seifert, H., Threlfall, W.: A Textbook ofTopology. Academic Press: New
York, 1980
[31] Rokhlin, V.A., Fuchs, D.B.: Beginner's Course in Topology. Geometric
Chapters. Nauka: Moscow, 1977. English trans!.: Springer-Verlag: Berlin
Heidelberg New York, 1984
References 319

[32] Novikov, S.P., Fomenko, A.T.: Basic Elements of Differential Geometry


and Topology. Nauka: Moscow, 1987. English transl.: Kluwer Academic
Publishers: Amsterdam, 1990
[33] Fomenko, A.T.: Variational Principles in Topology. Multidimensional
Minimal Surface Theory. Moscow: Nauka, 1982. English transl.: Kluwer
Academic Publishers: Amsterdam, 1990
[34] SchrOder, E.: DUrer Kunst und Geometrie. Akademie-Verlag: Berlin, 1980
[35] Milnor, lW.: Morse theory. Ann. Math. Studies 51 (1963)
[36] Mathematical Encyclopedia, vols. 1-4, 1977-1984 (Russian)
[37] Smale, S.: The Classification of Immersions of the Spheres in Euclidean
Spaces. Ann. Math., SeT. 2, 69, No.2 (1959) 327-344
[38] Boy, W.: liber die Curvatura integra und die Topologie geschlossener
Flachen. Dissertation. Gottingen, 1901. Math. Ann. 57 (1903) 151-184
[39] Francis, G.K., Morin, B.: Arnold Shapiro's eversion of the sphere. Math.
Intell. 2 (1979) 200--203
[40] de Rham, G.: Varietes differentiables. Formes courants, formes har-
moniques. Publ. de l'lnst. Math. de l'Univ. de Nancago, Hermann, Paris
(1955)
[41] Birman, S., Hilden, M.: The homeomorphism problem for 8 3• Bull. Am.
Math. Soc. 79, No.5 (1973) 1006-1010
[42] Volodin, lA, Kuznetsov, V.E., Fomenko, A.T.: On the problem of al-
gorithmic recognition of a standard three-dimensional sphere. Usp. Mat.
Nauk 24, No.5 (1974) 71-168 (Russian)
[43] Homma, T., Ochiai, M., Takahashi, M.: An algorithm for recognizing 8 3
in 3-manifolds with Heegaard splittings of genus two. Osaka l Math. 17
(1980) 62~8
[44] Casler, B.G.: An embedding theorem for connected 3-manifolds with
boundary. Proc. Am. Math. Soc. 16, No.4 (1965) 559-566
[45] Raken, W.: Theorie der Normalflachen. Acta Math. 105 (1961) 245-375
[46] Matveev, S.V., Savvateev, V.V.: Three-dimensional manifolds with sim-
ple special splines. Colloq. Math. 32, F. 1 (1974) 83-97
[47] Matveev, S.V.: One way of defining 3-manifolds. Vestn. Mosk. Univ.,
Ser. I, Mat., Mekh., No.3 (1975) 11-20 (Russian)
[48] Matveev, S.V., Fomenko, A.T.: Constant-energy surfaces of Ramiltonian
systems, enumeration of three-dimensional manifolds in increasing order
of complexity, and computation of volumes of closed hyperbolic mani-
folds. Usp. Mat. Nauk 44, No.1 (1988) 5-22. English transl.: Russ. Math.
Surv. 43, No.1 (1988) 3-24
[49] Fomenko, A.T.: Symplectic Geometry. Methods and Applications. Mos-
cow State University Press: Moscow, 1988. English transl.: Gordon and
Breach: London, 1988
320 References

[50] Kozlov, V.V.: Integrabiiity and nonintegrability in Hamiltonian mechan-


ics. Usp. Mat. Nauk 38, No.1 (1983) 3-fJ7 (Russian)
[51] Arnold, V.I., Varchenko, A.N., Gusein-Zade, S.M.: Singularities of Dif-
ferentiable Mappings. Classification of Critical Points, Caustics and Wave
Fronts. Nauka: Moscow, 1982 (Russian)
[52] Koblitz, N.: Introduction to Elliptic Curves and Modular Forms. Grad.
Texts Math. 97. Springer-Verlag: New York Berlin Heidelberg, 1984
[53] Koblitz, N.: P-adic Numbers, p-adic Analysis and Zeta-Functions. Grad.
Texts Math. 58. Springer-Verlag: New York Berlin Heidelberg, 1977
[54] Fomenko, AT.: Morse theory of integrable Hamiltonian systems. Dokl.
Akad. Nauk SSSR 287, No.5 (1986). English trans!.: SOy. Math. Dokl.
33, No.2 (1986) 502-506
[55] Fomenko, A.T.: The topology of surfaces of constant energy in integrable
Hamiltonian systems, and the obstructions to integrability. Izv. Akad.
Nauk SSSR, No.6 (1986). English trans!.: Math. USSR, Izv. 29, No.3
(1987) 629-fJ58
[56] Kharlamov, M.P.: The Integral Mapping Method in Rigid-Body Dynamics
Problems. Moscow State University Press: Moscow, 1980, 17-18 (Rus-
sian)
[57] Kharlamov, M.P.: Symmetry in gyroscopic systems. Mekh. Tverd. Tela
15 (1983) 87-93 (Russian)
[58] Pogosyan, T.I., Kharlamov, M.P.: Domains of possible motion in some
mechanical systems. Prikl. Mat. Mekh. 45, No.4 (1981) 605-fJ1O (Rus-
sian)
[59] Tatarinov, Ya;V.: Global View of Rigid-Body Dynamics. The Description
of Configuration. Vestn. Mosk. Univ., Ser. I, Mat., Mekh., No.4 (1978)
101-109 (Russian)
[60] Fomenko, AT.: Topological invariants of Liouville-integrable Hamilto-
nian systems. Funkts. Anal. Prilozh. 22, No.4 (1988) 38-51 (Russian)
[61] Fomenko, AT.: The symplectic topology of completely integrable Hamil-
tonian·systems. Usp. Mat. Nauk 44, No. 1(1989) 14S-c173. English trans!.:
Russ. Math. Surv. 44, No.1 (1989) 181-219
[62] Fomenko, A.T.: Integrability and Nonintegrability in Geometry and Me-
chanics. Kluwer Academic Publishers: Amsterdam, 1988
[63] Seifert, H., Threlfall, W.: Variationsrechnung im Gr06en. Hamburger
Math. Einzelschr. 24. Teubner: Leipzig, 1932
[64] Brailov, A.V., Fomenko, AT.: Topology of integral manifolds of com-
pletely integrable Hamiltonian systems. Mat. Sb. 133, No. 3 (1987)
375-385. English trans!.: Math. USSR, Sb. 62, No.2 (1989) 373-383
References 321

[65] Fomenko, A.T., Zieschang, H.: On typical topological properties of in-


tegrable Hamiltonian systems. Izv. Akad. Nauk SSSR 52, No.2 (1988)
378-407. English trans!.: Math. USSR, Izv. 32, No.2 (1989) 385-412
[66] Matveev, S.V., Burmistrova, A.B.: The Structure of S-Functions on Ori-
entable 3-Manifolds. Theses ofthe XI-th All-Union School on the Theory
of Operators in Functional Spaces. Vo!. 1. Chelyabinsk, 1986 (Russian)
[67] Morgan, J.: Non-Singular Morse-Smale flows on 3-dimensional mani-
folds. Topology 18, No.1 (1979),41-54 .
[68] Matveev, S.V., Fomenko, A.T., Sharko, V.V.: Round Morse functions
and isoenergy surfaces of integrable Hamiltonian systems. Mat. Sb. 135,
No.3 (1988) 325-345 (Russian)
[69] Fomenko, A.T., Zieschang, H.: On the topology ofthe three-dimensional
manifolds arising in Hamiltonian mechanics. Dokl. Akad. Nauk SSSR
294, No.2 (1987). English trans!.: SOY. Math., Dokl. 35, No.2 (1987)
529-534
[70] Bolsinov, A.V., Matveev, S.V., Fomenko, A.T.: Topological classification
of integrable Hamiltonian systems with two degrees of freedom. A list of
low-complexity systems. Usp. mat. Nauk 45, No.2 (1990) 49-77. English
trans!.: Russ. Math. Surv. 45, No.2 (1990), 59-94
[71] Waldhausen, F.: Eine Klasse von 3-dimensionalen Mannigfaltigkeiten. I.
Invent. Math. 3, No.' 4 (1967) 308-333
[72] Zieschang, H.: Finite Groups of Mapping Classes of Surfaces. Lect. Notes
Math. 875. Springer-Verlag: Berlin Heidelberg New York, 1981
[73] Fomenko, A.T., Trofimov, V.V.: Integrable Systems on Lie Algebras and
Symmetric Spaces. Gordon and Breach: London, 1988
[74] Thurston, W.P.: Three-dimensional manifolds, Kleinian groups and hy-
perbolic goemetry. Bull. Am. Math. Soc. 6 (1982) 357-381
[75] Thurston, W.: Hyperbolic structures of 3-manifolds. Ann. Math. 124
(1986) 203-246
[76] Weeks, J.R.: Hyperbolic Structures on Three-Manifolds. A dissertation
presented to the Faculty of Princeton Univ., Dept. of Math., 1985
[77] Fomenko, A.T.: Symmetries of soap films. Compo Math. with App!. 12B,
Nos. 3/4 (1986) 825-834
[78] Dao Trong Thi, Fomenko, A.T.: Minimal Surfaces, Stratified Multivari-
folds, and the Plateau Problem. Nauka: Moskau, 1987. English trans!.:
Trans!. Math. Monogr. 84, 1990.
[79] Pogorelov, A.V.: Differential Geometry. Nauk:a: Moscow, 1974. English
trans!.: Noordhoff: Groningen, 1967
[80] Aleksandrov, A.D.: Intrinsic Geometric of a Convex Surface. Gostekhisz-
dat: Moscow Leningrad, 1948
322 References

[81] Pogorelov, A.V.: Extrinsic Geometry of Convex Surfaces, Nauka,


Moscow, 1969. English trans!.: Trans!. Math. Monogr. 35, 1973
[82] Pontryagin, L.S.: Topological Groups. Nauka: Moscow, 1973. English
edition: Gordon and Breach: New York London Paris, 1966
[83] Kirby, R.: Stable homeomorphisms and the annulus conjecture. Ann.
Math. 89 (1969) 575-582
[84] Chernavsky, A.V.: Local contractibility of the group of homeomorphisms
of manifolds. Mat. Sb. 79 (1969) 307-356 (Russian)
[85] Novikov, S.P.: On manifolds with a free abelian fundamental group
and their applications. (Pontryagin classes, smoothness, multi-dimenional
knots). Izv. Akad. Nauk SSSR, Ser. Mat. 30, No.1 (1966) 207-246 (Rus-
sian)
[86] Efremovich, VA, Chemavsky, A.V.: Elements ofTopology. Yaros!. Uni-
versity Press: Yaroslavl, 1977 (Russian)
[87] Birman, J.: Braids, Links and Mapping Class Groups. Ann. Math. Stud.
82. Princeton University Press: Princeton, 1983
[88] Freedman, M.: The topology of four-dimensional manifolds. J. Diff.
Geom. 17 (1982) 357-452
[89] Hatcher, A., Thurston, W.: A presentation for the mapping class group of
a closed orientable surface. Topology 19 (1980) 221-237
[90] Zeeman, E.C.: Catastrophe Theory. Selected papers 1972-1977. Addison-
Wesley: Reading, Mass., 1977
[91] Zenkin, A.A.: The extension of the Wiferich theorem to the case of natural
summands. Dokl. Akad. Nauk SSSR 264, No.2 (1982) 282-285 (Russian)
[92] Zenkin, A.A.: Generalization of the Hilbert-Waring theorem. Vestn.
Mosk. Univ., Ser. I, Mat., Mekh., No.2 (1983) 11-19 (Russian)
[93] Tatarinov, Ya.V.: Lectures on Classical Mechanics. Moscow State Uni-
versity Press: Moscow, 1984 (Russian)
[94] Tatarinov, Ya.v.: The construction of compact invariant manifolds, other
than tori, in one integrable nonholonomic system. Usp. Mat. Nauk 40,
No.5 (1985) (Russian)
[95] Okuneva, G.G.: Movement of a rigid body with a fixed point under non-
holonomic constraints in a Newton field. Mekh. Tverd. Tela 18 (1986)
40-43 (Russian)
[96] Okuneva, G.G.: Qualitative research of integrable version of the Suslov
non-holonomic problem. Vestn. Mosk. Gos. Univ., Ser. Mat., Mekh., No.
5 (1987) 59-64 (Russian)
[97] Coxeter, H.S.M.: Introduction to Geometry. John Wiley and Sons Inc.:
New York London, 1961
[98] Bourbaki, N.: Elements de mathematique. Groups et algebres de Lie.
Hermann: Paris, 1971, 1972
References 323

[99] Coxeter, H.S.M.: Discrete groups generated by reflections. Ann. Math.


35 (1936) 58~21
[100] Andreev, E.M.: On the intersection of the planes of faces of polyhedra
with acute angles. Mat. Zametki 8, No.4 (1970) 521-527 (Russian)
[101] Makarov, B.S.: On extension of discrete groups of the Lobachevskian
plane up to spatial ones. Uch. Zap. Kishin. Univ. 82 (1965) 56-59 (Rus-
sian)
[102] Makarov, V.S.: On one division class of Lobachevskian space. Dokl.
Akad. Nauk SSSR 161 (1965) 277-278 (Russian)
[103] Makarov, V.S.: On one class of discrete groups of the Lobachevskian
space with an infinite fundamental region of finite measure. Dokl. Akad.
Nauk SSSR 167 (1966) 30-33 (Russian)
[104] Makarov, V.S.: On one class of two-dimensional Fedorov groups. Izv.
Akad. Nauk SSSR, Ser. Mat. 31 (1967) 531-542 (Russian)
[105] Andreev, E.M.: On convex polyhedra in Lobachevskian spaces. Mat. Sb.
81, No.4 (1970) 445-478 (Russian)
[106] Andreev, E.M.: On finite-volume convex polyhedra in Lobachevskian
spaces. Mat. Sb. 83, No.2 (1970) 256-260 (Russian)
[107] Vinberg, E.B.: On the groups of unities of some quadratic forms. Mat.
Sb. 87 (1972) 18-36 (Russian)
[108] Shaikheev, M.K.: Mapping subgroups in Bianchi groups. Vopr. Teor.
Grupp i Gomologicheskoj Algebry (Probl. of Group Theory and Homo-
logical Algebra) No.2 (1987)
[109] Lanner, F.: On complexes with transitive groups of automorphisms.
Comm. Sem. Math. Univ. Lund 11 (1950) 1-71
[110] Kaplinskaya, I.M.: On discrete groups generated by reflections in faces
of simplicial prisms in Lobachevskian spaces. Mat. Zametki 15 (1974)
159-164 (Russian)
[Ill] Vinberg, E.B.: The absence of crystallographic reflection groups in
Lobachevskian spaces of high dimension. Tr. Mosk. Mat. O-va 47 (1984)
68-102 (Russian)
[112] Prokhorov, M.N.: The absence of discrete reflection groups with a non-
compact fundamental finite-volume polyhedron in a Lobachevskian space
of high dimension. Izv. Akad. Nauk SSSR, Ser. Mat. 50, No. 2 (1986)
413-424 (Russian)
[113] Bugaenko, V.O.: On automorphism groups of unimodular hyperbolic qua-
dratic forms over the ring 12:[r3 + 1/2]. Vestn. Mosk. Gos. Univ., Ser. 1,
Mat., Mekh. No.5 (1984) 6-12 (Russian)
[114] Vinberg, E.B., Kaplinskaya, LM.: On the groups 0 18 ,1(12:) and 0 9,1(12:).
Dokl. Akad. Nauk SSSR 238 (1978) 1273-1274 (Russian)
324 References

[115] Vinberg, E.B.: Discrete groups generated by reflections in Lobachevskian


spaces. Mat. Sb. 72 (1967) 471-488 (Russian)
[116] Bugaenko, V.O.: On reflective unimodular hyperbolic quadratic fonns.
Vopr. Teor. Grupp i Gomologicheskoj Algebry (Probl. of Group Theory
and Homological Algebra) No.4 (1987)
[117] Vinberg, E.B.: Hyperbolic reflection groups. Usp. Mat. Nauk 40, No.1
(1985) 29-66 (Russian)

Acknowledgements

We would like to thank the original publishers of A.T.Fomenko' s pictures for granting
permission to reprint them here. The numbers following the sources correspond to the
numbering of the pictures in this book.
A.T. Fomenko: Computers and Mathematics with Applications, Visual and
Hidden Symmetry in Geometry, Symmetry 2, Vol. 17, No. 1-8, Pergamon Press,
I. Margittai, 1989: 1.1.25, 1.2.20,2.1.19,3.2.41,4.1.6,4.2.9
1. Stoyanov: Counterexamples in Probability, John Wiley & Sons, 1989: 1.3.12

You might also like