OceanofPDF - Com Topology - Richard Earl
OceanofPDF - Com Topology - Richard Earl
Available soon:
KOREA Michael J. Seth
TRIGONOMETRY Glen Van Brummelen
THE SUN Philip Judge
AERIAL WARFARE Frank Ledwidge
RENEWABLE ENERGY Nick Jelley
TOPOLOGY
A Very Short Introduction
Great Clarendon Street, Oxford, OX2 6DP, United Kingdom
Oxford University Press is a department of the University of Oxford. It furthers the University’s
objective of excellence in research, scholarship, and education by publishing worldwide. Oxford is a
registered trade mark of Oxford University Press in the UK and in certain other countries
© Richard Earl 2019
The moral rights of the author have been asserted
First edition published in 2019
Impression: 1
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, without the prior permission in writing of Oxford
University Press, or as expressly permitted by law, by licence or under terms agreed with the
appropriate reprographics rights organization. Enquiries concerning reproduction outside the scope of
the above should be sent to the Rights Department, Oxford University Press, at the address above
You must not circulate this work in any other form and you must impose this same condition on any
acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
Library of Congress Control Number: 2019949429
ISBN 978–0–19–883268–3
ebook ISBN 978–0–19–256899–1
Printed in Great Britain by Ashford Colour Press Ltd, Gosport, Hampshire
Links to third party websites are provided by Oxford in good faith and for information only. Oxford
disclaims any responsibility for the materials contained in any third party website referenced in this
work.
In memory of
Dan Lunn.
Friend, colleague, tutor.
Contents
Acknowledgements
List of illustrations
1 What is topology?
2 Making surfaces
3 Thinking continuously
5 Flavours of topology
Epilogue
Appendix
Historical timeline
Further reading
Index
Acknowledgements
2 Examples of polyhedra
11 The torus
12 Making a cylinder
13 Making a torus
23 Examples of graphs
32 Visualizing convergence
As you read this, passengers the world over are travelling on metro (or
subway or underground) trains. There are around 60 billion individual
journeys made annually on such metro systems. But whether this be in
Tokyo, London, São Paolo, New York, Shanghai, Paris, Cairo, Moscow,
those travellers are perusing maps for their journeys that are crucially
different from maps in atlases or seen on geography classroom walls.
Foremost in the minds of those passengers are the connections they need to
make—getting out at the right station and changing to the correct new line.
They are not interested in whether the map’s left–right lines do indeed run
west–east, or whether they really did make a right angle turn when they
changed lines, as depicted on the metro map.
The oldest metro network in the world is the London Underground. When
first produced, the underground maps superimposed the different train lines
onto an actual (geographically accurate) map of London, as shown in
Figure 1(a). A first version of the current map was designed by Harry Beck
in 1931 as in Figure 1(b). Beck’s map, and the current underground map,
are not wrong. Rather they transparently show information important to
travellers—for example, the various connections between lines and the
number of stops between stations. It is an early example of a topological
map and demonstrates the different focus of topology—which is all about
shape, connection, relative position—compared with that of geometry (or
geography) which is about more rigid notions such as distance, angle, and
area.
1. London underground maps (a) Geographically accurate 1908 map, (b) Beck’s topological
1931 map.
Topology is now a major area of modern mathematics, so you may be
surprised to learn that an appreciation of topology came late in the history
of mathematics. The word topology—meaning ‘the study of place’—wasn’t
even coined until 1836. (‘Geometry’ by comparison is an ancient Greek
word and ‘algebra’ is an Arabic word, with its mathematical meaning
dating back to the 9th century.) Just why this was the case is not a simple
question to address, though we will see some aspects of topology developed
as mathematicians sought to put their subject on a more rigorous footing.
Topology is a highly visual subject that lends itself to an informal treatment
and this book will give you a sense of topology’s ideas and its technical
vocabulary.
A topologist’s alphabet
As a first example, to convey how differently topologists and geometers see
objects, consider what capital letters a topologist would deem to be the
‘same’. Using the sans serif font, the four letters
EFTY
are all topologically the same. They are not congruent, meaning that none
of the letters can be picked up, and rotated or reflected, and then put down
as one of the other letters. But I hope you can envisage, if allowed to bend,
stretch, or shrink the letters, how any of them might be transformed into one
of the others.
C G I J L M N S U V W Z.
Topologically all of these are equivalent to a line segment and it’s not hard
to imagine how each might be formed by bending and stretching a suitably
mutable letter I. So hopefully you’re convinced that the letters in the second
list are all homeomorphic to one another, but what makes this second
collection topologically different from the first list?
Note, for each letter E, F, T, Y, that every point lies on a distorted bit of line
with one exception. In each of these letters there is a single point that might
be described as a T-junction. These T-junctions are highlighted below.
One way in which these T-junctions are special is that, if removed, the
remainder of the letter is disconnected into three parts; the removal of any
other point would leave just two parts remaining. In whatever ways we
might bend and deform an E the deformed version would still include a
single T-junction. As none of the second set C … Z has such a T-junction
then none of them can be a deformed version of an E (or an F, T, or Y ).
This gives a genuine sense of how mathematicians resolve the question: are
two shapes the same topologically? This either amounts to finding some
means of continuously deforming one into the other, or involves finding
some topological invariant of one that does not apply to the other. The
word invariant is used in different contexts in mathematics: for example, if
you shuffle a pack of cards, there will still remain fifty-two cards afterwards
and four suits, these are invariants; but the top card may have changed and
the jack of clubs may no longer come before the eight of diamonds, and so
such facts aren’t invariants of a shuffle. A topological invariant is
something immutable about a shape, no matter how we stretch and deform
it. In the above example we used the presence of a T-junction as our
topological invariant. You might note that an E includes four right angles
whilst an F contains only three. The presence of four right angles is a
geometric invariant and so shows that E and F are not congruent (i.e. not
geometrically the same), but—working topologically—we are permitted to
unbend these right angles and so right angles are not important from a
topological point of view. Rather they’re mutable aspects of a shape and not
topological invariants.
You might want to take a moment thinking about what makes an A different
from a P or O different from Q. In fact, the O introduces an important
topological invariant that separates it from both I and E. The shape of the O
is different as it makes a loop. Technically O is not simply connected, a
topic we will discuss more in Chapter 5.
Euler’s formula
One of the first topological results was due to Leonhard Euler (pronounced
‘oil-er’), a titan of 18th-century mathematics and one of the most prolific
mathematicians ever; his formula dates to around 1750. The result relates—
at first glance—to polyhedra, three-dimensional objects such as cubes and
pyramids (Figure 2). It is also so fundamental—a straightforward
observation at least—that it is surprising ancient Greek mathematicians
missed it.
2. Examples of polyhedra (a) A cube, (b) A Square-based pyramid.
Looking at the cube, we can see that it is made up of vertices (the corners
of the cube—the singular is ‘vertex’), these vertices being connected by
edges and that these edges then bound (square) faces. For the cube the
number of vertices V equals 8, there are E = 12 edges and F = 6 faces. For
the (square-based) pyramid we have V = 5, E = 8, and F = 5. No pattern
may be evident immediately but if we include the four other so-called
Platonic solids—tetrahedron, octahedron, dodecahedron, icosahedron
(Figure 4)—and other familiar polyhedra, we create Table 1.
(We shall see soon that the truncated icosahedron is familiar to us—just not
by that name!)
For all the geometry that the ancient Greeks knew, it seems striking that this
pattern eluded them, but we will prove now—or more honestly sketch a
proof of—Euler’s formula which states, for a polyhedron with V vertices,
E edges, and F faces, that
In the proof our aim will be to begin with a polyhedron and manipulate it in
certain ways—for example, we might remove or subdivide faces—but in all
cases we will carefully track the effect our manipulation has (if any) on the
number . If, after such manipulations, we arrive at a simplified
situation where we know what equals, and we know the effects
our manipulations had on that number, then we may be able to work
backwards to find what was originally.
We begin then with a polyhedron, and first remove one of the faces. This
has the effect of reducing by 1 as F has decreased by 1. Now
that the polyhedron has a missing face—effectively the polyhedron has
been punctured—it can be flattened into the plane, taking care that all the
vertices, edges, faces present on the punctured polyhedron remain and are
connected in the plane in the same manner they were on the punctured
polyhedron. For example, if we removed one face from a cube and flattened
the remaining cube then we would have something like Figure 3(a).
3. Manipulations of a cube to find its Euler number (a) A flattened, punctured cube, (b) With
flattened faces triangulated, (c) Removing a triangle, (d) Removing a triangle.
We now remove the triangles one at a time. For example, if we remove the
bottom triangle from Figure 3(b) to make Figure 3(c), then we remove one
edge and one face and, by the same reasoning as before, this has no overall
effect on .
Eventually only a single triangle will remain, having removed all others. A
triangle has a single face, three vertices, and three edges, so that
equals . This is the value of that we
finish with. The only manipulation that ever changed was that
very first removal of a face which reduced it by one; initially then it was the
case that
It’s also worth mentioning that René Descartes had, over a century before
Euler, demonstrated a theorem for polyhedra that is equivalent to Euler’s
formula; his theorem was in terms of ‘angular defects’ at vertices. All of a
sudden we are back in the geometrical world and it’s less than clear that
there is a genuinely new subject, an importantly different way of
mathematical thinking, that Euler’s fingertips were brushing against.
Euler’s formulation encourages appreciation of the result as something a
little new—in Euler’s terminology, rather than Descartes’s, it’s much clearer
that the connection of the vertices, edges, and faces is what counts, but
historically we are still a long time from a deeper appreciation of topology
as a fundamental mode of mathematical thinking.
If we are seeking to be rigorous here, we should really point out that the
previous calculations show that there are at most five possible pairs of
values that m, n can take. Those calculations limit the possibilities, but do
not necessarily mean that there is a Platonic solid for each of these cases,
nor preclude there being more than one Platonic solid for permitted m and n
—it might be that there are two different Platonic solids with three
pentagons meeting at each vertex. Listed in Table 2 are the five Platonic
solids, and so we can see that there is at least one solid for permitted m, n.
And it’s not hard to appreciate why there can be at most one. In the case
where , then three squares meet at each vertex; seemingly this
only tells us something about parts of the solid, but if we follow this recipe
of attaching three squares at each vertex then there is only one way to
progress building up the solid—it’s not clear that this recipe will actually
lead to a complete solid, but it does show that there can be at most one
Platonic solid for each allowed m, n.
Footballs
In 2017 the mathematics popularizer Matt Parker began a petition seeking
to get road signs to football stadia corrected.
You may not have noticed the inaccuracy of such signs in the past, perhaps
being happy just to know you’re travelling the right way for the game. But
it’s clear (Figure 5) that the sign’s football does not resemble the actual
football that Matt is carrying. A football’s surface is made from pentagons
and hexagons and the everyday football is more formally known as a
truncated icosahedron. (It can be created from an icosahedron by planing
down, around each vertex, the five edges meeting there. If we plane down
one-third of each of those five edges, we create a new pentagonal face and
continued planing eventually shrinks all the triangular faces to hexagons.) I
think though the irksome principle for Matt was not that the sign’s football
was badly drawn, it was in fact impossibly drawn.
In fact, Euler’s formula shows us how many pentagons and hexagons there
are on a football. Recalling how to truncate an icosahedron we see there are
as many pentagons as original vertices (12) and hexagons as original faces
(20), but Euler’s formula can show this is the only way to construct such a
football. Say a football has P pentagonal faces and H hexagonal faces.
Then, before gluing these together, we have 5P + 6H unglued edges and the
same number of unglued vertices. Looking at Matt’s ball we can see that (i)
two unglued edges are needed to make an edge on the football and (ii) three
unglued vertices make a vertex with (iii) two hexagons and one pentagon
meeting at a vertex; from (iii) we see there are twice as many ‘unglued’
vertices collectively on the hexagons as on the pentagons. So
If we put these values into Euler’s formula we find that
which simplifies and rearranges to P =12 and the equation 10P = 6H yields
H =20. This, then, is the only way to make a football if we follow the rules
(i), (ii), (iii).
Graph theory
Let’s change tack a little and consider the following problem. The square
PQRS in Figure 6 has diagonally opposite vertices P and R, Q and S. If we
were to draw curves from P to R, and from Q to S, curves which remain
within the square as in Figure 6, then surely those curves would have to
cross at least once. (In Figure 6 there are three intersections.) This seems
obvious—and is true—but how would you go about proving this?
At first glance, this problem might seem quite removed from the polyhedra
we were just discussing. However, Figure 6 doesn’t look that different from
Figures 3(a)–(d). We have vertices (P, Q, R, S and any points where the
curves PR and QS meet), edges running between these vertices (though
admittedly they’re now curved), and we have faces bounded by those edges.
It was crucial to the proof of Euler’s formula that for each
of Figures 3(a)–(d). If we also include the outside region as a face—
essentially the one removed so we could flatten the polyhedron—then we
arrive back at Euler’s formula . (By this reckoning ,
, in Figure 6.)
Some graphs are planar, meaning that they can be drawn in the plane
without their edges crossing (at points that aren’t vertices). The two graphs
K5 and K3,3 in Figure 7 are importantly not planar. The complete graph on 5
vertices, denoted K5, has a single edge between each pair of the 5 vertices
making 10 edges. You might think that K5 is planar as it’s drawn in Figure
7(a)—the point is that, so drawn, many of the edges’ crossings don’t occur
at vertices and to deem these crossings as vertices would mean we were no
longer considering K5 which has only 5 vertices. If we could properly draw
K5 in the plane there would be 10 triangular faces v1v2v3, v1v2v4, through to
v3v4v5. We’d then have that equals and
so K5 is not planar.
7. Two non-planar graphs (a) The complete graph K5, (b) The complete bipartite graph K3,3.
Nasty surprises
Euler arrived at his formula a century before the word topology was coined.
His formula is characteristic of a visual side of topology naturally aligned
with geometry. But topology, as a subject, would develop along various
themes and in particular had an important role in the foundational work
mathematicians were doing around the start of the 20th century. As I hinted
earlier, topology’s rise may have been hampered by a traditional mindset
that some of its questions had obvious answers. For example Camille
Jordan, as late as the 19th century, proved the following: a curve in the
plane, which does not cross itself and which finishes back where it began—
a curve which we would now call a Jordan curve—splits the plane into
two regions, the technical phrase for these regions being connected
components. One of these regions is bounded, the inside, and the other is
unbounded, the outside, and this is the Jordan curve theorem. Earlier
mathematicians would have happily thought this obvious and the first
rigorous proof didn’t appear until 1887. You may agree with those earlier
mathematicians that the result can be safely assumed. Maybe even the
Pollock-like Jordan curve in Figure 8(a) does not sway your view of the
intuitiveness of the result.
8. Complicated Jordan curves (a) A more complicated Jordan curve, (b) A Jordan curve with
positive area.
In Figure 8(b) is the Knopp–Osgood curve which, for all its fractal-like
appearance, is a Jordan curve. Astonishingly it has a positive area—that is
the curve itself has positive area, we’re not referring to some region that it
bounds. Would you have said a moment ago that it’s obvious that curves
can’t themselves have area?
You shouldn’t worry too much in the sense that most things those early
mathematicians thought to be true turned out to be true, once properly
understood and qualified, but mathematicians towards the end of the 19th
century were getting nervous about the rules and assumptions that
mathematics relied on.
A Flatland mindset
The novella Flatland: A Romance of Many Dimensions, written in 1884 by
Edwin Abbott, is a satire on Victorian mores. The narrator is ‘A Square’, an
inhabitant of Flatland, a planar world having just two dimensions. The
culture of Flatland and the logistics of living in two dimensions are fully
described, implicitly highlighting some of the narrow-mindedness of
Victorian culture—for example, women are one- rather than two-
dimensional beings. The story doesn’t explicitly discuss topology, but in its
description of worlds with different dimensions and implications for the
inhabitants, it provides a useful metaphor for understanding certain aspects
of topology.
The knottedness of the trefoil says something about its position in 3D. In
fact, all knots are homeomorphic to a circle. To better appreciate this, you
might imagine life as an ant living on either the circle or trefoil. As the ant
moves around either the unknot or trefoil it has a sense of being on a loop,
but the ant has no notion of whether it is living on a knot. It is only by being
able to view things from outside the two loops, and looking on from a
position in the ambient space, that we are able to recognize one loop as
knotted as compared with the other. This Flatland mindset will prove useful
again later when we meet subspaces.
Topology would advance on various fronts in the 19th and 20th centuries.
In particular, Bernhard Riemann would early on show the usefulness of a
‘topological mindset’, introducing Riemann surfaces into the study of
polynomial equations and demonstrating some deep connections between
topology and many other areas of mathematics.
Chapter 2
Making surfaces
For Figure 11(a)’s ‘polyhedron’, a count of vertices, edges, and faces shows
that , , , giving which seems to
disprove Euler’s formula. We are left with a few alternatives: either the
object in Figure 11(a) should not be considered a polyhedron, or we need to
restrict Euler’s formula to a certain type of polyhedron, or we need to adapt
and generalize Euler’s formula into a version that remains true for a broader
family of polyhedra.
11. The torus (a) A polyhedron with one hole, (b) A torus.
The most obvious issue with this new ‘polyhedron’ is the hole through its
middle. This is not immediately reason enough to exclude it as a
polyhedron, but this shape, once a face is removed, does not leave a
remainder that can be flattened into the plane, making our earlier proof
invalid. We need either to restrict Euler’s formula to polyhedra without
holes, or we need to work out the correct values for polyhedra
with holes.
12. Making a cylinder (a) A square with identified edges, (b) Making the cylinder, (c) A
cylinder.
13. Making a torus, (a) Torus with v and e1, e2 drawn, (b) Square with gluing instructions, (c)
K3,3 on the torus.
Note that the four corners of the original square—denoted v1, v2, v3, v4—
have all been glued together to make a single point v on the torus. Similarly,
the edges e1 and e3 have been glued together to form a circle going around
the outside of the torus and e2 and e4 have been glued together to form a
different circle going through the hole of the torus, these two circles
meeting at the point v.
As an aside, the graph K3,3, which we saw can’t be drawn in the plane
(Figure 7(b)), can be drawn on the torus (Figure 13(c)). The reason K3,3 can
be drawn on the torus is because a torus has a lower Euler number of 0 and,
arguing as before, it can then be shown that F = 3. If you look carefully at
Figure 13(c) you will see that there are indeed three faces (quadrilaterals
v1w2v2w1 and v3w3v2w2 and a single hexagonal face v1w2v3w1v2w3).
How does all this help with determining Euler numbers? With a single
square and gluing directions, we have been able to make an object with the
shape of a torus and this is certainly a conciser description of a torus than
Figure 11(a) for which , , . But is this glued square
enough to work out the Euler number of a torus? The answer is yes if we’re
careful when thinking about just how the original vertices and edges glue
together. On the original square there was just one face—the square’s
interior—four unglued edges (labelled e1, e2, e3, e4), and four unglued
vertices (labelled v1, v2, v3, v4). However, once we’ve followed the gluing
directions, those four edges have become the two circular edges on the torus
and the four vertices have become one point v, as in Figure 13(a). And there
is still one ‘face’ on the torus—the square’s interior has been stretched to
become all of the torus except those circles and v. So when we use this
glued square to calculate the Euler number of the torus we get the answer
If you prefer Figure 13(a) showing the torus with the four glued vertices
becoming one, and the four edges become two loops on the torus, then
Figure 13(b) will only appear as an unfinished DIY job. But the single
surface obtained from gluing the triangle, pentagon, and square in Figure
14, following all the gluing directions a, b, … f according to the arrows,
might reasonably start stretching your visualization skills. But we can still
work out the Euler number of this chimera and seek to understand just what
surface we are looking at. This time there are three faces (the triangle,
pentagon, and square) and the twelve unglued edges of the polygons make
six glued edges a, b, c, … f on the surface. How many vertices will we
ultimately have? There were twelve unglued vertices originally but various
of these get glued together as we make the surface. For example, v1 and v5
are glued together as they are both at the rear end of the edge marked a. In
fact, we can chase around these gluings to see just how many vertices we
have:
14. Gluing instructions for a triangle, pentagon, and square.
So eight different (unglued) vertices v1, v2, v4, v5, v7, v9, v11, v12 all get
glued together as a single vertex on the surface. In a similar fashion we can
see that the remaining four vertices v3, v6, v10, v8 get glued together (in that
order). So once made, the surface has 2 vertices, 6 edges, and 3 faces giving
an Euler number of . Just what surface have
we made?
Looking at these so-called six subdivisions, only two of these are in fact
permissible, 15(c) and 15(d). In 15(a), 15(e), 15(f), there is a face that is not
a distorted polygon; neither the whole sphere (15(a)) nor the punctured
cummerbund (15(e)) nor the twice-punctured sphere (15(f)) are
topologically the same as a polygon and so not permissible faces. In 15(b)
and 15(e), the edges do not begin and end in a vertex. 15(e) is deliberately
given to show that the correct Euler number can be incorrectly calculated.
Looking back at the torus in Figure 13(a) and the surface in Figure 14, we
calculated their Euler numbers using subdivisions consistent with the above
three rules. Therefore, we correctly calculated their Euler numbers as 0 and
–1 respectively.
Connected sums
Given two closed surfaces S1 and S2, then we can create their connected
sum S1#S2. This is a way to glue surfaces together and a useful means of
making new surfaces from the few we have so far met. Say S1 and S2 each
has a subdivision that includes a ‘triangular’ face bounded by three edges.
(The inverted commas here hint that, this being topology, the faces may not
be that recognizably triangular in terms of having straight edges.) The
connected sum S1#S2 is then created by removing these two triangular
faces, so making two holes in the surfaces, and then gluing the two surfaces
together along the boundaries of the holes, pairing up the three vertices and
three edges with those on the boundary of the second removed face as, for
example, in Figure 16.
16. Connected sums with tori (a) One torus with a “triangle” missing, (b) A torus with two
holes as a connected sum.
Helpfully there is a formula for the Euler number of S1#S2. In making the
connected sum, we remove two triangular faces, the six different vertices on
these triangles are glued to make three vertices on the connected sum, and
likewise six edges are glued to make three. So the total number of faces has
gone down by 2 and the total numbers of edges and vertices have each gone
down by 3. As V and F are added in the formula for the Euler number, and
E is subtracted, overall we have
Or, if you prefer a more careful algebraic proof, say the original subdivision
of S1 has V1 vertices, E1 edges, and F1 faces and define V2, E2, F2 similarly
for S2. The number of vertices V#, edges E#, and faces F# on the connected
sum is given by
Finally
Thinking in terms of connected sums helps us work out the Euler numbers
of some more complicated surfaces. We know that a torus has an Euler
number of 0. The connected sum is a torus with two holes (Figure
16(b)) and we see
and similarly the torus with three holes, , has Euler number
In fact, we can see that every time we make a connected sum with the
surface gains one more hole and the Euler number reduces by 2. So the
torus with g holes—which can be considered as , the connected sum of
g copies of the torus —has Euler number
The number g of holes in the surface is called the genus of the surface.
One-sided surfaces
At this point, we still can’t identify the peculiar surface from Figure 14
which has an Euler number of –1. So far we’ve only constructed surfaces
with even Euler numbers and –1 is odd. In fact, with the tori , we’ve
only met half the story and half of the closed surfaces. Recall how in Figure
12 we made a cylinder by gluing two edges of a square. We could, instead,
have glued those two sides using reversed arrows (Figure 17(a)),
introducing a single twist. So the points near v2 on e1 are glued to the points
near v4 on e3 and those near v1 on e1 are glued to the points near v3 on e3.
This would have created a Möbius strip, named after August Möbius who
discovered it in 1858.
17. The Möbius strip (a) A square with identified edges, (b) Runners on a Möbius strip.
19. The Klein bottle and projective plane (a) A Klein bottle, (b) 3D depiction of a Klein Bottle,
(c) A projective plane.
There is a subtle problem with the Klein bottle in Figure 19(b). When we
take the cylinder back into itself, some single points in space actually
represent two distinct points on the Klein bottle. So this image is not a
proper representation or embedding of the Klein bottle in 3D. In fact, it is
impossible to construct a Klein bottle in 3D without such self-intersections
as occur where the cylinder cuts back into itself. The relevant result
demonstrating this impossibility can be viewed as a generalization of the
Jordan curve theorem. That theorem concerned embedding circles in the
plane with a Jordan curve having an inside and an outside. In a like manner
when a closed surface is embedded in 3D, the surface again divides the
remaining space into an inside and an outside and so the closed surface
must be orientable. As the Klein bottle is non-orientable, it cannot be
embedded in 3D.
However, the Klein bottle can be embedded in 4D and this isn’t too hard to
imagine if we treat the fourth dimension as time. The Klein bottle is two-
dimensional (as surfaces are) and so from this 4D viewpoint it is important
to consider the Klein bottle as only existing for an instant, a certain ‘now’;
for it to have a past or future would give it a third dimension. So when
faced with bringing the cylinder back into itself—which would normally
cause self-intersections—we can instead move that bit of cylinder gradually
into the future (the fourth dimension), where the remainder of the Klein
bottle doesn’t exist and then, once the cylinder has passed through the space
its present self occupies, we can gradually bring that bit of the cylinder back
into the present. The self-intersections no longer occur, as the distinct points
of the Klein bottle that became merged in Figure 19(b) instead sit in the
same point of space but crucially at different times.
We can also determine the Euler number of the Klein bottle, again being
careful to note how edges and vertices are glued together. The square is our
only face; e1 and e3 are glued together, as are e2 and e4, making two rather
than four edges; finally v1 is glued to v2 which is glued to v4 which is glued
to v3 and so we have just one vertex, giving ,
the Euler number of the Klein bottle. Unfortunately, 0 is also the Euler
number of the torus, so any hope we might have had that the Euler number
alone is information enough to recognize the shape of a surface was
simplistic. The torus and Klein bottle are different surfaces—the former is
orientable (two-sided), the latter not—and yet they both have the same
Euler number.
Another important non-orientable surface, which can be formed from
gluing a square’s edges together, is the projective plane ℙ. In Figure 19(c)
we assign e2 and e4 reverse arrows (in contrast to 19(a)). The surface
formed is non-orientable, as it again contains a Möbius strip (the shaded
region), and we can calculate the Euler number as before: again and
but this time v1 and v3 are glued together and separately v2 and v4
are glued, so that . Hence ℙ has Euler number
.
It turns out that the Euler number goes a long way to separating out the
different surfaces, but we have seen that this cannot be the whole story as
the torus and Klein bottle have the same Euler number whilst being
different surfaces—the first is orientable, the second not. The only missing
ingredient in the classification is that notion of orientability.
So the first half of the classification theorem for two-sided surfaces states:
And the second half of the classification theorem for one-sided surfaces
states:
Overall then, the classification theorem says that if we know the Euler
number of a closed surface and whether it is one- or two-sided, then we
know its topological shape. If you were wondering, where the Klein bottle
is on this list, we know its Euler number to be 0 and we know it to be one-
sided. The only surface in the classification matching these facts is the
surface ℙ#ℙ and this is topologically the same as the Klein bottle.
We might create a yet more complicated connected sum such as
which at first glance is not on our list. This surface is
one-sided and its Euler number equals
so topologically it’s the same surface as ℙ#7. And at long last we are able to
identify the surface we formed in Figure 14. That surface had Euler number
–1 and so the surface is ℙ#ℙ#ℙ, this being the only surface on our list with
that Euler number.
Complex numbers
Surfaces are a natural two-dimensional extension of one-dimensional
curves which mathematicians had long been interested in but, historically,
surfaces and their topology became of particular importance because of the
work of 19th-century mathematicians, most notably Bernhard Riemann.
has no real numbers as solutions. If you take a positive number x then its
square x2 is also positive (and so cannot equal –1); if you take a negative
number then its square is also positive; finally . So there are no
solutions. If you prefer a more pictorial approach then you might draw the
graphs of and , and the fact that these graphs don’t meet
(Figure 20(a)) is again another way of showing that no number x solves the
equation . Basically the problem is that negative numbers don’t
have real square roots.
20. The real line and complex plane (a) Graphs of y = x2 and y = –1, (b) The real line, (c) The
complex plane.
And there the story might have ended except those Renaissance
mathematicians found good reasons to ‘imagine’ that does have
solutions, denoting a solution as i. This may seem somewhat ludicrous at
first, but around 1530 a method was found for solving degree 3 equations.
One problem was that this method necessitated calculations with square
roots of negative numbers, even when all the equation’s solutions were real
numbers. The worth of the number i became truly apparent with the proof
of the fundamental theorem of algebra in 1799 by Carl Gauss. This
theorem shows all the solutions of any polynomial equation have the form
where a and b are real numbers. For example, the number
solves the equation
In the same way that real numbers are commonly represented on the real
line (Figure 20(b)) the complex numbers can be represented as a plane, the
complex plane (Figure 20(c)). A complex number such as can then
naturally be identified with the point (1, 2) as shown. The real numbers
occupy the horizontal axis—denoted ‘Re’—and the vertical axis ‘Im’ is
called the imaginary axis.
Complex numbers have a rich theory of their own which, for
mathematicians at least, is reason enough to warrant their study. You may,
though, be surprised to find that quantum theory, the physical theory that
successfully models subatomic physics, is naturally described using the
language of complex numbers and so physicists, chemists, and engineers all
need to be well versed in the use of complex numbers.
Riemann surfaces
The introduction of complex numbers led to a much richer theory
connecting algebra and geometry. In Figure 20(a) we see that the curves
and don’t meet; if they did meet at a point (x,y) in the real
xy-plane then we’d have and no such x exists. But using complex
numbers they do intersect at two points, at and at
. The fact that a degree 2 curve and a degree 1 curve meet
in points in this case is not entirely coincidental. More generally
it is the case that, if properly counted, a degree m curve and a degree n
curve intersect in points. Multiple contacts need to be counted
properly—so a line tangentially meeting the curve would count as a
double contact, so that there are still intersections. The final
finesse, when counting intersections, is to include points at infinity. For
example, two parallel lines—each degree 1 curves—are still deemed to
meet at a point at infinity so that there is intersection as expected.
The curve sits in the real xy-plane as a curved version of the x-axis
(Figure 20(a)); the curve and axis are topologically the same with a
homeomorphism between the two just pushing each point (x, 0) up to the
point (x, x2). If we include also the curve’s point at infinity bringing
together the curve’s ‘ends’ then the curve topologically becomes a circle.
When using complex numbers, the curve sits in complex xy-space
as a curved version of the x-axis which, remember, is itself a two-
dimensional complex plane. When we include the point at infinity this
brings together this curved plane as a sphere. (This is the reverse process of
puncturing a sphere to get the plane that we met earlier in the proof of
Euler’s formula.)
21. Visualizing Riemann surfaces (a) Graph of y2 = x(x – 1)(x – 2), (b) Graph of y2 = x(x – 1)2,
(c) A pinched torus.
where g is the genus of the Riemann surface and d is the degree of the
curve’s equation. Remembering the examples we have met, note that
for gives , a sphere, and for
gives , a torus. For curves with singular points, the formula can be
generalized including a correction term for each singularity, as shown by
Max Noether in 1884.
Given just one sentence for the task, many topologists might choose to
describe their subject as the study of continuity. The word ‘continuous’
appeared a few times in Chapter 1 but it was left to the reader’s intuition as
to quite what the word entailed. In many ways this reflects how
mathematicians used to regard continuity—historically it was just
considered evident what was meant by ‘continuous’ and as many (but not
all) of the results about continuous functions that are ‘obvious’ also happen
to be true, relatively little effort was spent providing further clarity. A
rigorous definition of continuity did not appear until the 19th century.
From Figure 22(b) we can see that the car stops at t3—where the speed s(t)
becomes zero—and after t4 increases to the speed limit. The distance
travelled d(t) in Figure 22(a) is a continuous function of time t. Historically
this would have been understood as meaning its graph could be drawn
without taking pen from paper, but we will seek to provide a fuller
understanding. But the acceleration function a(t) is not continuous because
of the jumps in the graph in Figure 22(c). The times t1, t2, … t6 of
discontinuity in the acceleration relate to the driver’s foot coming off the
accelerator, being put on the brake, coming off the brake, and then the
pattern repeats again.
Functions
The idea of a function is a central one to mathematics, though this has only
been true since around the 17th century. Once Descartes and Fermat
independently introduced the idea of Cartesian coordinates x and y to
describe position in a plane, a curve could just as easily be described by an
equation as by its geometry. For example, the curve is a parabola, a
curve the ancient Greeks would have investigated solely using geometry. A
sketch of this curve is given in Figure 23(a)—the curve’s equation gives a
rule for plotting, above each point (x, 0) of the x-axis, a point (x, x2). Note
how certain algebraic properties of the function are represented in the shape
and position of the curve—as x2 ⩾ 0 for all x, the curve lies entirely on or
above the x-axis; as , the curve is symmetric about the y-axis.
23. Examples of graphs (a) Graph of , (b) Graph of y =.
For some functions, we might naturally have to limit the allowed inputs—or
we might choose to do so anyway. For example, if then we at
least need to ensure that x is non-zero as division by zero is meaningless;
for the function , then we cannot permit x to be negative, as no
real number has a negative square (Figure 23(b)).
There is still quite a bit of subtlety needed to fully capture what this means.
In our example, g(x) has a jump of 1 from output values near 2 (just before
) to output values near 1 (just after ). The size of that jump was
unimportant, the presence of any jump at all was sufficient. And the notion
of ‘nearby inputs’ should not be interpreted as several inputs that are in
some sense close to 1; rather we mean there are inputs x arbitrarily close to
1 such that g(x) is noticeably different from g(1). Necessarily this means
that we are talking about infinitely many such inputs x, not just several x.
By way of example, it is enough to note that:
This rigorously shows that g(x) is discontinuous at . The sequence of
inputs 0.9, 0.99, 0.999, 0.9999, … gets arbitrarily close to 1. What this
means is: however demanding ‘nearby to 1’ is required to be, there are
inputs from this sequence that are at least that close.
h(x) is noticeably different from h(0) for some inputs x arbitrarily near
to 0.
Near the function h(x) is varying crazily. From the graph we can see
that there are inputs x, arbitrarily close to 0, where whilst we have
. It now seems clear by our emerging sense of continuity that h(x)
is discontinuous at .
An example in detail
We still need to be careful turning these nascent thoughts into a rigorous
definition. We’ll consider in detail the function which is
continuous for all inputs x. If f(x) is continuous at an input then—
based on our previous thoughts—we need that
f(x) is not noticeably different from f(a) for all inputs x suitably near to
a.
if inputs x and 0 differ by less than d then outputs x2 and 0 differ by less
than e.
As the outputs here are closer to one another than the inputs are—that is, as
x2 is closer to 0 than x is to 0—then we can just choose d to equal e. So if
inputs differ by e or less so do the outputs. We have then shown that
is continuous at .
A rigorous definition
Putting all this thinking together gives us a rigorous definition of continuity.
I’d suggest reading the definition and seeking to understand how this means
that the function in Figure 25(a) is continuous and the one in Figure 25(b)
isn’t, but if you find the generality of the definition and the technical level
of the language difficult then move on to the next section on the properties
of continuous functions. And be reassured, as it took generations of
mathematicians to finally get this definition right, and current and past
generations of mathematics undergraduates still wrestle with proofs
involving this definition in their analysis courses.
25. The rigorous definition of continuous and discontinuous (a) A continuous function, (b) A
discontinuous function.
Formally, then, a function with real inputs x and real outputs f(x) is
continuous at an input if:
such that the difference between the outputs f(x) and f(a) is less than e
Note how this captures there being inputs x arbitrarily close to a (as they
can be found within any distance d of a) for which the inputs f(x) and f(a)
are ‘noticeably different’ (here meaning differing by more than e). In Figure
25(b) if we choose e to be smaller than the jump in the output that occurs at
then outputs f(a + d) will be greater than f(a) + e and so outside the
permitted range—no matter how small we make d.
which is sketched in Figure 26(b). Note that its only outputs are –1 and 1,
so that there are definitely no solutions to . But to prove the
intermediate value theorem, we would need to show that a function f(x)
satisfying the theorem’s requirements takes the value zero somewhere—
importantly we would know very little specifically about f(x) save that it is
continuous, begins negatively, and finishes positively, so any approach to a
proof would have to be similarly general. Visualizing drawing the graph
from beneath to above the x-axis, it seems as though there would have to be
a first time that we cross the x-axis and this is indeed the case. That input
value c where we first cross would be
26. The intermediate value theorem (a) Intermediate value theorem, (b) A discontinuous
function.
But writing down this definition of c doesn’t itself constitute a proof and
carefully proving the intermediate value theorem is well beyond the aims of
this book. The above line is the proof’s starting point, we have a candidate
input c where we think the function is zero, but it remains to carefully show
that f(c) = 0. The intermediate value theorem was first proved by
Bolzano in 1817.
27. The blancmange function (a) y = f1(x), (b) y = f2(x), (c) y = f3(x), (d) Blancmange function.
To conclude, recall the earlier comment that a function is the whole package
of domain, codomain, and the assignment rule. We cannot simply say
whether x2 is a bounded function. The assignment x2 is part of an
unbounded function when the domain and codomain are both the real line,
but when we restrict the domain to the interval a ⩽ x ⩽ b then the
assignment x2 yields a bounded function. This pre-empts somewhat a more
detailed discussion for Chapter 4: what is it about the domain a ⩽ x ⩽ b
that means continuous functions are bounded on that domain or satisfy the
intermediate value theorem? The answer to the first question is that the
interval is compact and to the second question is that the interval is
connected. We will see in Chapter 4 what these terms mean more generally.
Chapter 4
The plane and other spaces
More on functions
In Chapter 3 we discussed the continuity of functions that take one
numerical input and produce one numerical output. Most functions are not
of such a simple form. As you read this, the density of matter in the room
around you is a function of three spatial coordinates (needed to describe a
point of the room you’re in) and one coordinate describing time. If you are
reading the paperback version, that density function takes a roughly
constant value (the density of paper) at points in the book you’re reading
but that value changes (discontinuously) at the book’s edges and takes on a
new value (the density of air) at points outside the book. If you’re reading
this outside, the wind’s velocity is an output with three components
measuring to what extent the wind is blowing ahead/behind, up/down,
left/right; each of these three components is again a function of one
temporal and three spatial inputs. We’d expect wind velocity to be a
continuous function, even if it may sometimes change quite quickly. To
have you thinking a little harder, is it reasonable to say that the distance a
car has travelled is a continuous function of its speed? This really is a subtle
question as neither the input nor output are numbers, but rather functions of
time, with input the speed function s(t) and output the distance function d(t)
(Figure 22). If you have some sense of what the question is asking—will
my journey to work tomorrow be much the same as my journey today if I
keep to much the same speed during my journey?—then your intuition is
doing well. Even then, there are important details to be filled in, namely
describing in each case just what ‘much the same’ means.
Thoughts on distance
Consider the important notions needed to define continuity in Chapter 3.
Loosely put, continuity requires that we can constrain the difference in
outputs by suitably constraining the difference in inputs. We will need a
more general notion of the difference between other types of input and
output: for example, what is the ‘difference’ between two points of the
plane?
Given two real numbers, x and y, we denote the difference between them as
|x–y|, and this is just another expression for the distance between x and y as
points on the real line (Figure 20(b)). Likewise, given two points (x1, y1)
and (x2, y2) in the plane, we might take the ‘difference’ between them to
mean the straight-line distance between them (Figure 28(a)). Pythagoras’
Theorem tells us this distance equals
28. Visualizing different metrics (a) Straight line distance, (b) Taxicab distance, (c) Distance
between functions.
so that the points (1,1) and (3,2) in Figure 28(a) are a distance
apart. But in some circumstances, you might decide that
the straight-line distance is not the best way to describe distance
realistically. For example, a taxi driver in Manhattan would be constrained
by New York’s grid of streets and so would need to take a journey as in
Figure 28(b), travelling along perpendicular streets and avenues. That
distance is given by the formula
For the taxi driver, the points (1,1) and (3,2) are a distance apart,
as the taxi may not go along the straight path through Manhattan’s
skyscrapers.
the maximum distance between f(x) and g(x) as we range over all inputs
x,
as highlighted by the two dotted lines. But another reasonable notion of the
distance between the functions might be
Again, what makes the better definition of distance between two functions
may depend on what the functions signify.
The first property says that distances can’t be negative and the second
implies that the distance between two distinct points is positive. The third
says that the distance going from one point to another is the same as the
distance coming back.
As we have seen there are different choices of metrics on the same set—for
example straight-line or taxicab on the plane—and two different metrics
might have quite different views of whether two points are close to one
another or not. So it is quite possible that the same function between two
sets might be continuous when using certain metrics, and discontinuous for
a different choice of metrics.
Based on the graphs, you may think that f(x, y) and h(x, y) are continuous,
whilst g(x, y) is discontinuous—all of which is true. You may further
suspect that g(x, y) is discontinuous at the points (0, y) on the line ,
which is again correct. The jump in output g(x, y) across the line , as
we move between the two rules that define the function, is precisely why
g(x, y) is discontinuous there. h(x, y) is also defined by two separate rules,
one for inputs where , one for when . The graph
of h(x, y), when is the bowl-shaped part of Figure 29(c). As
we move towards the boundary of that rule’s application, the circle
, then gets ever closer to equalling 1
which is the rule for h(x, y) outside of the disc. So whilst h(x, y) is defined
by two rules, the rule for nicely hands over to the second rule
for without any jump in the output. This is why h(x, y) is
continuous.
As before there are nice algebraic results guaranteeing that if f (x, y) and
g(x, y) are continuous functions from the plane to the real numbers then so
are the functions
Our choice of metric does matter for F, which evaluates an input f (x) at
. Continuity means being able to constrain the outputs by constraining
the inputs: must two functions be close at 0 if the functions f (x) and g(x) are
close? The answer is yes if we’re using the maximum-distance-apart metric;
the difference between the functions at the single input 0 can be no more
than the greatest distance between the functions when considering all
inputs. However, we cannot constrain the difference between the functions
at 0 by constraining the area between the functions (Figure 29(e)). A
function g(x) with a tall, but very thin spike around would produce a
large difference between f (0) and g(0) whilst f (x) and g(x) would be close
in terms of the area-between metric.
And we can now properly ask and answer the question that started this
chapter: is the distance a car has travelled, d(t), a continuous function of its
speed, s(t)? (See Figure 22.) Any journey to work taking time T or less can
be represented by a continuous function d(t) on the interval 0 ⩽ t ⩽ T. Is
the function that takes input s(t) to output d(t) continuous? If two cars have
speeds s1(t) and s2(t) that never differ by more than S then the two cars will
never be more than distance ST apart on their journeys. So using maximum-
distance-apart, we can constrain the distance between two journeys d1(t)
and d2(t) by constraining the distance between their speeds s1(t) and s2(t).
We have just shown d(t) is a continuous function of s(t) when using the
maximum-distance-apart metric. Using a little calculus, a similar argument
can be made for the area-between metric.
However, if we were using the taxicab distance in the plane the ball B(a, r)
looks different. The taxicab distance between two points and
equals and so the points within
(taxicab) distance r of a0 are those satisfying
Some algebraic manipulation shows that these points form the interior of a
diamond (Figure 30(a)).
It is possible to fit such a diamond into any disc of the plane, and vice versa
it is possible to fit a disc into any given diamond (Figure 30(b)). This is
what it means for two metrics to be equivalent. More precisely, two
different metrics d1 and d2 are said to be equivalent if any ball B1(a, r)
contains some ball B2(a, s) and any ball B2(a, R) contains some ball B1(a,
S). The subscripts here refer to which metric is being used and the smaller
ball’s radii s and S will usually be different from the radii r and R of the
original balls.
for any points a and b, and where the subscripts SL and T signify the metric.
So BT(a, r) is contained in BSL(a, r) which is contained in , for
any point a and radius r (Figure 30(b)).
The important point here is that two different but equivalent metrics lead to
the same functions being continuous. Now we can rewrite the definition of
continuity in terms of open balls: a function f(x), with inputs from M and
outputs in N, is continuous at input if
for any positive e there is some positive d such that f sends B(a, d) into
B(f(a), e).
The preimage f –1(U) consists of all elements of M that f sends into U. For
example, if and U is the interval 0 < x < 2 then the pre-
image f–1(U) is the interval –1 < x < 1 as these are precisely those x that
satisfy the inequality .
That different metrics can lead to the same continuous functions and that
continuity can be rephrased in terms of open sets, making no mention of
metrics, suggest that the open sets, more so than metrics, are crucial to
continuity. However, if we are going to take this approach—starting with
open sets rather than metrics—we need to decide what properties open sets
need to have; so far we have only defined open sets in terms of metrics. The
two main properties of open sets are:
Given a collection of sets their union consists of those elements that are in
one or more of those sets; the intersection of those sets consists of those
elements that are in every one of those sets (see Figure 31(c)). The union of
any collection of open sets is open, but in general the intersection of an
infinite collection of open sets need not be open. Consider the interval –r <
x < r where r is a positive number. If we consider all such sets for positive r,
then the only element in each set is 0. So the intersection of these sets is
{0}, the set with sole element 0, an intersection which is not open.
Note that M, by itself, is ‘just’ a set, not a metric space, with elements
having no sense of being close to one another; the collection of sets is an
effort to separate out M without necessarily going so far as introducing a
metric. Given a metric space then its open sets form a topology, but not all
topologies arise from metrics.
The smallest topology on M just includes the empty set and the set M.
This is known as the trivial topology—it does not separate out the elements
of M at all and the only continuous functions on M are the constant ones.
The largest topology is the discrete topology in which case every set is in ,
including single points; it introduces extreme separation, placing each
element of the set away from others and all functions on M are continuous.
In practice, important topologies are somewhere between these two
extremes. Many important topologies arise from metrics, but there are
important ones that do not, such as the Zariski topology, important in
algebraic geometry which studies sets defined by polynomial equations.
Working with topological spaces, rather than metric spaces, does more than
just generalize further the ideas of continuity. Many proofs in topology
appear cleaner, with the logic of the proof and the rules of being a
topological space meshing together much more naturally.
The ellipsis ‘. . .’ at the end of each list denotes that the list goes on forever.
A sequence is commonly denoted as x1, x2, x3, … and the terms in these
sequences can each be described by giving a formula for the nth term xn:
Only the first of these sequences converges; the terms of the sequence are
steadily getting smaller and I hope it’s not surprising that this sequence
converges to 0 which is known as the sequence’s limit. The second
sequence does not converge to any limit; some of its terms though—like the
even terms (second, fourth, sixth, etc.) which are all 1—do converge but
overall the sequence does not. The final sequence does not converge either
and in fact no selection of its terms converges either.
The very word ‘converge’ suggests that the terms of the sequence get closer
and closer to some limit. Rigorously, a sequence xn, in a metric space M,
converges to a limit a in M if any ball B(a, r) contains a tail of the
sequence; this means that from some term onwards, all remaining terms of
the sequence are inside B(a, r).
Using this definition, we can see why the sequence converges to
0. The ball B(0, r) contains all terms xn of the sequence where n > 1/r,
which is a tail of the sequence. In Figure 32(a), a sequence spirals in to its
limit (0, 0). The pictured dashed disc B((0, 0), r) contains x10 and every
term afterwards; we would need to go farther down the sequence to find a
tail contained within a smaller open ball. If a sequence converges to a limit
in a metric space, then the sequence converges to the same limit if we use
any equivalent metric.
32. Visualizing convergence (a) A sequence converging in the plane, (b) Neither open nor
closed set.
So, an open set is one which contains none of its boundary points, and a
closed set is one which contains all its boundary points. Clearly most sets
fall into neither of these categories and will contain some but not all of their
boundary points; it’s important to realize that being open and being closed
are not opposites of one another. In Figure 32(b) is a set which is neither
open nor closed: a disc with its upper circumference included but the lower
half omitted. The boundary point a is contained within the set so the set is
not open—or equally no open ball B(a, r) around a is contained within the
set. And the boundary point b isn’t contained within the set, so the set is not
closed—or equally there is a sequence of points in the set that converges to
b despite b not being in the set. And some sets can be both open and closed,
for example the whole plane is both an open and closed set of the plane—
there are no boundary points, so all and none of them are simultaneously in
the set.
Subspaces
Any set S in a metric space M inherits the structure of a metric space, as two
points of S are also points of M and we may assign them the same distance
as before when they were considered as points of M. For example, if M was
the set of cities in the United States, and S is the set of Californian cities,
then it’s natural to think of Los Angeles and San Francisco as being the
same distance apart whether they’re being considered as American cities or
Californian cities. The set S naturally becomes a metric space, in its own
right, called a subspace of M.
Subtleties arise when we consider the open sets of and continuous functions
on a subspace. It may help to remember the Flatland mindset mentioned in
Chapter 1; we need to imagine life as an inhabitant of the subspace S, as if
Californians somehow cannot see the rest of the USA.
For example, let M be the real line and let S be the union of the intervals 0
⩽ x ⩽ 2 and 3 < x ⩽ 4 (Figure 33(a)). What are the open balls B(1, 1) and
B(2, 1) as open balls in S? By definition, B(1, 1) is the set of points in S
within distance 1 of the point 1; this is the interval 0 < x < 2 which agrees
with what B(1, 1) is in the real line M. However B(2, 1) is the interval 1 < x
⩽ 2 as an open ball of S; this is the set of points in S that are within distance
1 of 2. This is different from the situation with the real line, and is yet more
surprising when we remember open balls are open sets. The point 2 appears
incongruous as points near it to the right are not included but, from the
Flatland mindset of the subspace S, those points are essentially invisible as
they are not in S. So, the interval 1 < x ⩽ 2 is indeed an open set in S.
33. Relating to a disconnected subspace (a) Open balls in a union of two intervals, (b) Graph of
f(x) on S.
For the real line, plane, and higher dimensional equivalents, the Heine–
Borel theorem states that the compact sets are precisely the closed and
bounded sets. We have already defined what a closed set is, and a set is
bounded if it is contained in some ball B(0, R). So we can see that the real
line is not compact as it is not bounded and the interval 0 < x ⩽ 1 is not
compact as it is not closed, missing the boundary point of 0.
34. Relating to path-connectedness (a) A path-connected set, (b) A convex set, (c) Topologist’s
sine curve.
Path-connected metric spaces are connected, though there are some weird
spaces that are connected without being path-connected (Figure 34(c)). For
many regions, paths between points can be straightforwardly defined and a
space quickly shown to be path-connected and so connected. For example,
in an open disc, the plane, a half-plane, any two points are connected by a
straight line. These are all examples of convex sets, which means that given
any two points in the set then the line segment between them is also
contained in the set (Figure 34(b)) which is not true of the Pacman-like
shape (Figure 34(a)).
Topological invariants
In Chapter 1 we separated out which letters of the alphabet were
topologically the same—homeomorphic—or not. For two equivalent letters,
we described a way of deforming each into the other; for topologically
different ones, we needed to find a feature—a topological invariant, like the
T-junction in the E—that would remain a feature of the letter, even when
deformed. Compactness and connectedness are topological invariants and
so can be used to differentiate between spaces.
All the spaces are connected, but with a little imagination we can still make
use of connectedness to discern topological differences. A point of a
connected space which, when removed, disconnects the space is called a
cut point. We also noted in Chapter 1 that the T-junction in an E is the
special cut point which, if removed, breaks E into three connected
components; removing a different point leaves just two components
remaining.
Looking at the top row of Table 5, neither the circle nor the closed disc has
any cut points, but the figure 8 and interval do have; this means the former
two are not homeomorphic to the latter two. However, we can disconnect
the circle by removing two points and such is not true of the closed disc.
Finally, the middle point of the figure 8 is the only cut point whilst any
point of the interval, except the ends, is a cut point.
In the bottom row, the line and interval have cut points, but not the half-
plane nor plane. But 1 is not a cut point of the interval whilst every point of
the line is a cut point. All that remains to do is show the closed half-plane
and plane aren’t homeomorphic.
Taking the methods and theorems of continuity for single input, single
output functions and applying them to the more general settings of metric
and topological spaces proves to be a very powerful approach.
Consequently, properties such as compactness and connectedness are used
widely in mathematics and a classical theorem guaranteeing a real number
solution to an equation might in a modern setting show that there is a
continuous function which solves a differential equation. Seemingly these
are very different mathematical problems, but this abstract mode of thinking
helps mathematical ideas be applied in their fullest generality.
Chapter 5
Flavours of topology
In Chapter 2 we met the Euler number and saw that it, together with
knowing whether a surface is one- or two-sided, identifies the shape of a
closed surface. Further, the Euler number imposes global constraints about
what is possible on a closed surface as we’ll see in the next two sections.
Differential topology
The continuous functions form an important class of functions, but we also
saw in Chapter 3 that continuous functions can still be quite nasty; for
example, the blancmange function does not have a defined gradient at any
point. The smooth functions also form an important class. A function might
be continuous but still fail to be smooth by changing in a jerky fashion;
smooth functions by contrast have a defined gradient everywhere. For
example, the speed function s(t) in Figure 22(b) is continuous but doesn’t
have a defined gradient at the times t1, … t6. Before and after those times,
the gradient is clear but there is no defined gradient of s(t) when the car
accelerates or brakes suddenly. The study of functions that change smoothly
on surfaces lies within the field of differential topology and we will see
that the Euler number of a closed surface impacts on what properties
smooth functions on a closed surface can (or must) have.
An important first result about smooth functions is that the gradient of the
function is zero at any maximum or minimum. Note, in Figure 24(a), how
the gradient of sinx is zero at its extreme values whereas, in Figure 27(a),
the (continuous but not smooth) function f1(x) has no defined gradient at its
maxima. This result is known as Fermat’s theorem and these maxima and
minima are known as critical values and the corresponding inputs as
critical points.
But a function of two inputs can have a greater variety of critical points
than a function with just one input. When we have two inputs and one
output, the notion of gradient is a little less clear. Sketched in Figure 35 are
three surfaces exhibiting different types of critical point.
35. Examples of a minimum, maximum, and saddle point (a) , (b)
, (c) ,
Consider now the height function z for the torus drawn in Figure 36(a). The
height function has a maximum value at the top of the torus A, and a
minimum value achieved at the bottom of the torus D. In fact, as the torus is
compact, the boundedness theorem guarantees that any continuous function
on it must have at least one maximum and at least one minimum.
36. Critical points of functions on a torus and sphere (a) Height function on a torus, (b) z2 – y2
on the unit sphere.
There are two further critical points: C, at the bottom of the hole and B, at
the hole’s top. If we move through C, passing through the hole, then our
height is at its greatest as we pass through C; but if we slide down one side
of the hole and up the other side, then the lowest point of our journey is at
C. This critical point C is a saddle point and similar paths can be made
through B which include it as the highest or lowest point on the journey, so
B is also a saddle point.
• maxima at (0, 0, 1) and (0, 0, –1), the sphere’s north and south poles;
• minima at (0, 1, 0) and (0, –1, 0), two points lying on the sphere’s
equator;
• saddle points at (1, 0, 0) and (–1, 0, 0), two more points lying on the
sphere’s equator.
At the north pole (0, 0, 1), z is as large as it can be on the sphere, so any
move away means that z2 – y2 decreases and (0, 0, 1) is therefore a
maximum of f. The same argument applies at the south pole. Similarly, at
(0, 1, 0), y is as large as it can be, so any move away means that z2 – y2
increases and so (0, 1, 0) is a minimum, as is (0, –1, 0). Finally, at (1, 0, 0),
x is as large as it can be. As we move from (1, 0, 0) then x will decrease and
either y or z or both might increase, meaning that f may decrease or
increase. This means that (1, 0, 0), and likewise (–1, 0, 0) are saddle points
of f.
At this point, any relationship with topology probably seems unclear. The
height function on the torus had one maximum, one minimum, and two
saddle points; z2 – y2 on the unit sphere had two maxima, two minima, and
two saddle points. However, if we continued considering smooth functions
on the torus, in each case we would find that
(as with the height function where ) and for any smooth
function on the sphere we’d find that
In fact, we can see that the sphere (being the only closed surface with an
Euler number of 2 or more) is the only closed surface on which a smooth
function may have no saddle points. An example of such a function is the
height function on a sphere.
The Scottish physicist James Clerk Maxwell was one of the first to
appreciate such a relation amongst critical points in an 1870 paper On Hills
and Dales, but he was working solely with functions on the plane—and so
arrived at the number 1 (the Euler number of the plane or a punctured
sphere). Poincaré would generalize the result to closed surfaces, using
techniques that would now be considered Morse theory, after Marston
Morse who, from 1925 onwards, would prove a series of deep results
connecting the topology of spaces and the analysis of functions on those
spaces.
37. Vector fields on the sphere and torus (a) A hairy ball with no hair at poles, (b) A hairy-
everywhere torus.
The hair in this theorem is a metaphor for a tangent vector field which
might be easiest thought of as a fluid flowing on a surface; the hairy ball
theorem then states that for any fluid flow on a sphere, there will be one or
more points where the fluid is still and unmoving, the way in the eye of a
storm the wind is calm (Figure 37(a)). Such points, where the fluid is still,
are called singularities and there are various types of singularity that a
vector field can have. Three vector fields in the plane are shown in Figure
38, with the origin (0, 0) being the singular point for each.
38. Examples calculating indices of vector fields (a) , (b)
, (c) .
The velocity of the flow at the point (x, y) is given by a formula v(x, y).
Figure 38(a) has a source at (0, 0) with the fluid coming out of the origin;
Figures 38(b) and 38(c) are more complicated examples with fluid flowing
in and out of the origin along different lines.
So, in Figure 38(a) if we start at 12 o’clock the arrow points north and as
we move anti-clockwise the arrows move around to west at 9 o’clock, south
at 6 o’clock, east at 3 o’clock, and ultimately back to north. The arrows
themselves have gone once anti-clockwise around the compass (north–
west–south–east–north) and so the index is 1 in this case. In Figure 38(b) at
12 o’clock the arrows point east; as we move clockwise around the circle
they go from east to south (at 9 o’clock) to west to north to east—this
journey of east–south–west–north–east is a single journey once around the
compass clockwise and so the index in this case is –1. Finally Figure 38(c):
at 12 o’clock the arrow is pointing west, by 10.30 it’s pointing south, by 9
o’clock it’s pointing east, by 6 o’clock it’s already back to west, and by 12
o’clock the arrows have gone around one further time anti-clockwise before
returning to west. The arrows have been around twice in an anti-clockwise
fashion and the index in this case is 2.
(Indices is the plural of index.) If we look at the flow in Figure 37(a) then
we see that there is a source at the north pole which has index 1, and a sink
at the south pole which a check shows also has index 1 and then
adds up to the Euler number of the sphere. The hairy-everywhere torus in
Figure 37(b) has no singularities, so the sum of the indices is 0, the Euler
number of the torus.
Structures
Shortly we will move on to a discussion of algebraic topology which
associates with a surface (and other spaces) algebraic structures that capture
something of the essence of the surface’s topology. In Chapter 4 we met
compactness and connectedness, and these are important topological
invariants but, ultimately, they are also binary. A space can be compact or
not—there are no shades of grey here. It would be useful to have an
invariant that retained something subtler of the topological character of a
space. Before that we will need to consider the algebraic structures such
topological invariants might take the form of.
{0, 1, 2, 3, 4 . . .},
we might identify 0 as the smallest of these numbers, but then we’re already
recognizing the set’s implicit order, and we might recognize the numbers
can be added together but not subtracted, at least not if we want the result to
remain in the set (e.g. is in the set but is not).
One of the most common algebraic structures within mathematics is a
group. A group is a set G together with an operation * which, for two
inputs x and y from G, combines them into an output which we denote as x
* y and which is importantly also in G. The set G might be a set of numbers,
functions, geometric transformations or even a set of sets, and the operation
* might be addition, multiplication, composition of functions or some set
operation. But, to be a group, further rules must also apply: for G and * to
form a group we need:
1. for all x, y, z in G;
2. there is an element e in G such that for all x in G;
3. for any x in G there is an element x–1 in G such that
The element e in rule 2 is called the identity and the element x–1 which
combines with x to give e is called the inverse of x. Rule 1 is known as
associativity and the purpose of this rule is that any product like a * b * c *
d needs no further clarification—we’ll always arrive at the same answer no
matter in what order we carry out the three multiplications involved in this
product.
for any natural numbers x, y, and z, but fail rule 3, as there is no natural
number x–1 we can add to to get a sum of . We would like x–1
to be –2 but this is not a natural number, being negative. However, there are
many important sets and operations that do meet the three rules—the
subject group theory is a significant part of modern algebra—but as our
focus here is topology, we only mention a handful:
Algebraic topology
Our aim now is to associate with a space a group that captures something of
that space’s topological essence. The great French mathematician and
physicist Henri Poincaré had the idea to consider loops, paths in the space
that begin at a point and return to the same point. We might first try
appreciating Poincaré’s ideas on a torus. Our first problem is that there are
just too many loops; two different loops, such as l1 and l3 in Figure 39(a),
both go once around the hole in the torus in the same direction, but both
capture much the same about the shape of the torus. Secondly, Poincaré was
seeking to define a group and so he needed to find ways to combine loops
with some operation.
39. Loops on a torus (a) Loops on a torus, (b) As on a glued square, (c) Loops from a base
point.
Addressing this second point, we could combine loops by going around one
loop after going around the other; at least we could if the second loop began
where the first loop ended. So, first, we choose a fixed point of the torus, a
base point, and consider only loops that begin and end at the base point;
then, given two such loops l1 and l2, their product l2 * l1 is, reading from the
right, the path that starts at the base point, follows l1 and then l2, still
ultimately returning to the base point.
Would the set of loops based at a point, together with this operation *, make
a group? What, for example, would the identity loop e be, as described in
rule 2? This loop e would have to be such that for all loops l. This
means that the path l followed by e would have to be the same as just the
path l and so e would need to be the path that goes nowhere—e starts at the
base point and immediately finishes at the base point.
And what would the inverse l–1 of a loop l be, as described in rule 3? The
inverse l–1 of l must satisfy , knowing that e is the ‘don’t move’
path at the base point. This is problematic; even if we defined l–1 to be the
return journey of l, that is going around l in the opposite direction, e and l–
1* l would not be equal. To travel from London to Paris and back by the
same route is in no obvious way the same as just staying put in London.
Poincaré was able to resolve this issue with the same idea that he resolved
the first problem, that of there being too many loops. The two loops l1 and
l3 in Figure 39(a) seem to capture the same aspect of the topology of a
torus, but they’re different loops. However, either of them could be
continuously deformed into the other. The technical term for this is that the
two loops are homotopic. Any loop in the torus that is homotopic to l1 in
Figure 39(a) is a loop that goes once around the torus’s hole in the same
direction. But l2, a loop which goes once through the hole, is not homotopic
to either l1 or l3. In Figure 39(b) these loops are again depicted on a torus
represented as a glued square.
Why would it now be that following a path l, and returning along that same
path, is homotopic to making no journey at all? You might think of a
journey that did 99 per cent of the path l and returned from there, and then
one that did the same for 98 per cent, etc. Hopefully you now have the idea
that you can gradually do less and less of that return trip to Paris to the
point where you just stay in London—a kind of spectrum of increasingly
frustrating weekend breaks!
In Figure 39(c) a base point b is shown, and two loops l1 and l2, similar to
before but now based at b. A loop (based at b) that goes once around the
hole of the torus is homotopic to l1 and these loops are equal in the
fundamental group. A loop that goes once through the hole is homotopic to
l2. In fact it can be shown that any loop in the torus (based at b) is
homotopic to a journey around l1 some number of times followed by the
journey around l2 a certain number of times. What is being claimed here is
that knowing how many times a loop wraps around the torus’s hole and how
many times it wraps through the hole determines the essence (up to
homotopy) of a loop on the torus.
There is a subtle, implicit point here worth noting. The order in which the
loops wrap around or through the hole does not matter on the torus (but
does for other surfaces). This is because . This is easiest seen
in Figure 39(c). l1 is the loop running along the base (or top) of the square
and l2 is the loop running along the left (or right) side of the square. The
loop l2 * l1 is then the loop starting in the bottom left corner and going right
along the bottom and then up the right side, and l1 * l2 is the loop starting in
the bottom left corner and going up the left side and then right across the
top. We can see that l1 * l2 and l2 * l1 are homotopic—and so equal in the
fundamental group—by imagining the right-then-up path being
continuously dragged across the square until it becomes the up-then-right
path. The square’s diagonal, as in Figure 39(c), is a loop equal to either
product.
It may not now be that surprising to find that the fundamental group of a
circle is ℤ, the group of whole numbers under addition, as a loop in the
circle is essentially characterized by how many times it wraps around the
circle. Let’s remind ourselves of just what’s being claimed: a loop that goes
around the circle, starting and finishing at the same point, goes around the
circle a certain number of times (counted in an anti-clockwise sense).
Another, different loop may go around the loop the same number of times;
this second journey might have digressions here and there, back-and-forth,
but because the two journeys overall go around the circle the same number
of times then one journey can bit-by-bit be deformed into the second
journey; if the first journey gets behind/ahead of the second journey then it
can become increasingly hurried/delayed to rectify that. But a loop that
goes around the circle twice anti-clockwise could never be deformed into
one that goes once clockwise; these two journeys are essentially different—
that is, they’re not homotopic. These two journeys correspond to the
numbers 2 and –1 in the circle’s fundamental group. Any deformation of a
loop that goes twice around the circle will always be another loop that still
goes twice around the circle.
Finally, to say that the fundamental group of a circle is the group of whole
numbers under addition, means that loops around the circle combine in the
same way integers add. In the group ℤ of whole numbers, we know that the
numbers 2 and –1 add to give 1. In the fundamental group, 2 represents any
loop that goes twice anti-clockwise around the circle and –1 represents any
loop that goes once clockwise around the circle. If we combine such loops
using * then we get a loop that goes in total once anti-clockwise around the
circle and such a loop corresponds to 1 in the group of whole numbers. So
not only do the whole numbers correspond to loops around the circle, but
how whole numbers add corresponds to how loops combine using *.
As an example, consider when f is the function which wraps the circle twice
on to itself as shown in Figure 40(a). A point making angle θ with the
horizontal is sent to a point making angle 2θ with the horizontal axis—so
the arc from a to p would map to the arc from a to f(p), and the upper
semicircle from a to b would map to the whole circle. And if we had a loop
l in the circle that goes once (anti-clockwise) around the circle then f (l)
goes twice around the circle; if l goes twice around the circle then f (l) goes
four times around the circle. Loops that go n times around the circle are
what n represents in the fundamental group of the circle, and these loops are
sent to loops that go 2n times around the circle, which are what 2n
represents in the fundamental group. The function f wraps the circle twice
onto itself and naturally gives a corresponding function f for loops, with a
loop l that wraps n times around the circle being sent to a loop f(l) that
wraps 2n times around the circle.
40. Functions on the disc (a) f wrapping the circle twice around, (b) A function f from D to D,
(c) Defining the function g.
Say l1 and l2 are loops that wrap around the circle n1 and n2 times, so that l2
* l1 is a loop that wraps around the circle times. So f(l1), f(l2), f(l2 *
l1) are loops that wrap around the circle 2n1, 2n2, and times;
similarly f(l2)*f(l1) wraps around the circle times. As
then both f(l2 * l1) and f(l2)*f(l1) wrap around the circle an equal number of
times, meaning they are homotopic—that is
(H)
The theorem’s proof is not easy, perhaps involving the most conceptual
ideas in this text. First the proof is by contradiction which means that we
consider the possibility of there being a function f with no fixed points and
argue from that assumption to a position that is logically impossible. If a
function f has no fixed points, then for all x in D. Because of this
we can define the function g, from D to its circumference C as follows
(Figure 40(c)): as f(x) and x are different points, then we can draw a line
starting from f(x) and going through x which eventually meets the
circumference C at the point g(x). This way any point x in the disc D has
been sent to a point g(x) that lies on its circumference C. Note that if x is
itself on the circumference C, then .
• i has inputs in the whole numbers and an output of e, whatever the input,
so that this function is constant;
• g has a single input e and so some single output in the whole numbers;
• the composition g ° i, which has inputs and outputs in the whole
numbers.
Remember that the composition g ° i fixes all points of C and so any loop in
C is sent to the same loop by the corresponding function g ° i on loops. So,
in terms of the fundamental group of C, the function g ° i sends each whole
number n to itself. Instead we can consider the separate effects of doing the
corresponding function i first and then the corresponding function g. As
noted, i sends all the whole numbers to e, which g subsequently sends to
g(e). Note that this number g(e) is a single whole number not depending on
n.
Looked at one way, the function g ° i on loops sends all whole numbers n to
themselves, but considering the functions’ effects separately, every n is sent
to g(e); whatever this value g(e) is, it is a single value and so g ° i is a
constant function. This is the required contradiction—the function g ° i
cannot both send each n to n and also send each n to a single value g(e). So
a function f, without fixed points, cannot exist or else we’d be able to define
the function g and arrive at a contradiction.
The ideas of this proof are subtle, but the crux of it, and what makes such
methods powerful, is the following: fundamental groups (and other similar
algebraic invariants) are able to capture something of the topological
essence of a space; continuous functions between spaces lead to
algebraically nice functions (homomorphisms, that have the earlier property
H) between the spaces’ fundamental groups; if it can be shown no such
homomorphism exists, then no such continuous function existed.
where b1 captures there being 2g loops in the torus, one around each of the
g holes of the torus, and one in and through each hole. And the Betti
numbers of ℙ#k are
The Euler number of a space is defined in terms of its Betti numbers as the
alternating sum
so that the Euler number of equals and of ℙ#k
equals . That these Betti numbers are topological
invariants then means that the Euler number is a topological invariant.
For around the next forty years, topologists would take forward and make
rigorous Poincaré’s vision, making algebraic topology a major theme of
mathematics and fully appreciating the range and power of his ideas, with
James Alexander—more of him in Chapter 6—perhaps being most
prominent in that role. Arguably the last piece of Poincaré’s legacy came
with the proof in 2003 of the Poincaré conjecture by Grigory Perelman.
The conjecture states that every simply connected, compact three-
dimensional manifold is homeomorphic to the three-dimensional sphere
(this does not mean a solid ball in 3D, but rather a 3D spherical shell that
sits naturally in 4D). His proof, based on a strategy developed by Richard
Hamilton, used ideas of differential geometry far beyond even Poincaré’s
imagination at the time. However, its solution is more evidence for how the
great problems of mathematics have led to progress in highly novel
directions.
Chapter 6
Unknot or knot to be?
Describing knots
A knot is a smooth, simple, closed curve in 3D space. Being simple and
closed means the curve does not intersect itself except that its end returns to
its start. Basically a knot is a loop in 3D space and by requiring smoothness
we exclude some nasty, so-called wild, knots from our study. As I noted in
Chapter 1, all knots are topologically the same as a circle; what makes a
circle knotted—or not—is how that circle has been placed into 3D space.
So, two knots K1 and K2 are to be considered equivalent if we can start with
knot K1 and over a period of time continuously deform 3D space so that K1
becomes K2. If we set the time taken to deform the knots as a unit interval 0
⩽ t ⩽ 1, then by some time t in the middle of that interval we will have
deformed 3D space by some homeomorphism ht. An ambient isotopy is
then a continuous family of such ht, where at time we have yet to start
deforming so that and by the time the first knot has been
deformed into the second, that is .
The central problem of knot theory is then a classification theorem: when
are two knots equivalent—there is an ambient isotopy between them—or
how do we show that no such isotopy exists? This problem was first
identified by Maxwell in 1868, though his work went unpublished at the
time.
41. The unknot and trefoils (a) An unknot, (b) Left-handed Trefoil, (c) Right-handed Trefoil.
Knots only exist in 3D; any loop in 4D can be unknotted. Beginning from
the rightmost point of the unknot in Figure 41(a), and moving anti-
clockwise around the loop, we see that the two crossings both go over the
other part of the knot. More generally any loop that we could lay flat on the
table, at each crossing laying the loop over a previous part of the loop,
would lead to an unknot, as we could lift the loop off the table one crossing
at a time to make a circle. Imagine a genuine knot, with various under- and
over-crossings, but with us now permitted to move in 4D. Much as we
earlier used time as the fourth dimension to avoid the Klein bottle
intersecting itself, we could use this extra dimension to make any under-
crossing into an over-crossing; we could take the under-part of the knot
smoothly into the future, raise it above the over-part of the knot which only
exists in the present, and then smoothly return the knot from the future to
the present, now as an over-part. In this way all knots could be unknotted in
4D.
The trefoil is the simplest genuine knot and has three crossings. (You might
want to sketch for yourself loops with one or two crossings and convince
yourself these aren’t proper knots.) The trefoil though is complicated
enough to exhibit an important feature, that of chirality. The trefoil knots in
Figures 41(b) and 41(c) are mirror images of one another and aren’t isotopic
to one another. The trefoil is also an example of an alternating knot, a knot
where the crossings alternate between over- and under-crossings as we
travel around the knot. Alternating knots provide a somewhat simpler class
of knots which are well understood compared with knots generally.
Reidemeister moves
The unknot in Figure 41(a) can be quickly deformed into a circle by
untwisting it at the top, and doing similar to the twist at the bottom. Such a
move—a twist or untwist—is the first of three Reidemeister moves that
between them can be used to show any equivalent knots are indeed
equivalent. These moves are named after Kurt Reidemeister, who proved
this result in 1927, though the same result had been independently proved
by Alexander and Briggs in 1926, and in fact the moves were known much
earlier to Maxwell.
The first move then, R1, is a twist or untwist of a small part of the knot
(Figure 42). The second move R2 takes a small part of the knot and pokes it
under another part of the knot or undoes this. Finally, the third move R3 is
to slide a single crossing over a small part of the knot. Hopefully none of
these moves seems controversial and are evidently permitted manipulations
of the knot. The significant result is that, given any two equivalent knots,
repeated use of these three moves alone is sufficient to show that the two
knots are indeed equivalent.
43. The granny and reef knots (a) A granny knot, (b) A reef (or square) knot.
Note how this number of different prime knots spirals enormously as the
minimal crossing number increases. It is an open and active area of research
seeking to prove asymptotic estimates for how quickly the number of prime
knots grows as the crossing number becomes large. Figure 44 depicts these
prime knots up to a minimal crossing number of seven, all of which are
alternating; the simplest non-alternating prime knots have a minimal
crossing number of eight. In 1998 asymptotic estimates were demonstrated
for prime, alternating knots; the same result also showed that alternating
knots become increasingly rare, as a fraction of all knots, as the minimal
crossing number increases.
Recall that the elements of a fundamental group are loops based at a point,
with loops understood to be the same if one can be deformed into the other.
So, starting from and finishing at a point outside a knot, these different
loops might weave in and out of a knot helping to capture something of the
essence of the knot. And this is indeed the case, but unfortunately knot
groups remain very complicated objects. The knot group of the unknot (a
circle) is just the group of whole numbers, with a loop being entirely
characterized by how many times it wraps in and through the unknot. But
even the knot group of the trefoil is a difficult group to describe.
Wilhelm Wirtinger, around 1905, found a way to describe such knot groups
in general. In Figure 45 is drawn an oriented trefoil with the three unbroken
arcs between the crossings labelled as a1, a2, a3 and the five regions the
trefoil splits the plane into denoted R1 … R5. Imagine the base point b of the
knot group being outside the trefoil, and loops beginning at the base point,
weaving in and out of the knot, and returning back to the base point. We
will write l1 for a loop that goes down through R3 and returns back out of
R5, or equally down through R1 and returns back out of R4. This is a loop
from the base point b that ‘hooks’ a1 going in on the left of a1 and coming
back out on the right of a1. We denote by l2 and l3 similar loops that hook a2
and a3 in a left-to-right manner. Their inverses are loops that
hook a1, a2 and a3 in the reverse right-to-left direction. Any loop beginning
and ending in the base point b can be written as a string using the symbols
such as
45. The knot group of a trefoil.
a recipe for how the loop weaves in and out of the knot, with careful
attention as to whether the loop went in-left and out-right or vice versa. But
already we see that this can’t be the whole description of the knot group: in
the middle of the above ‘recipe’ is the expression which cancel out one
another (remembering back to Chapter 5 and the diminishing London–Paris
return trip). To do a loop and then do it in reverse is essentially the same as
not moving. So, we should omit all occurrences of expressions like
, etc.
However, there are other ‘relations’ between these loops l1, l2, l3 and there
is one such relation for each crossing of the knot. Consider a loop from b
that hooks the knot at the crossing between R1, R3, R4, R5; this is a loop that
goes down through R1 and back out through R5. This can be achieved as
l3l1, which means doing l1 first and l3 second; so we go in through R1 and
return out of R4 and then back in through R4 and out of R5. Or we could
manage the same by doing l1l2 which means doing l2 first and l1 second;
this takes us in through R1 and returns out of R3 and then down through R3
and out of R5. Either of these loops hooks the knot at the crossing point and
so it’s the case that l3l1 = l1l2.
Now if there were n crossings in a knot, then we’d have strings involving l1
… ln and in the knot group. Wirtinger’s theorem showed that the
knot group consists of all such strings, with two such strings describing the
same loop if one string can be turned into the other by cancelling terms like
or using rules, like l3l1 = l1l2, with one such rule coming from each
crossing.
All this is impressive, given its generality, but is not very tractable given
our aim is to describe simple invariants to separate out inequivalent knots.
It’s also frustrating to find that inequivalent knots, such as the reef knot and
the granny knot, can have the same knot groups, as do left- and right-
handed trefoils. However, it was shown in the late 1980s that the knot group
determines a prime knot up to mirror images: two prime knots with the
same knot group are isotopic knots, or each knot is isotopic to the mirror
image of the other.
The Alexander polynomial for the unknot is the constant polynomial 1 and
the Alexander polynomial of the trefoil T equals
So technically the Alexander polynomial is a polynomial in the variables x
and x–1. In his 1928 paper Alexander described an algorithm for calculating
his polynomial from an over- and under-crossings description of the knot
and that calculation, for the trefoil, is done in the Appendix. The algorithm
is a little technical but ultimately uses mathematics taught in schools and
colleges.
The Alexander polynomial also deals well with composite knots, having the
nice algebraic property
for two knots K and L, with K#L denoting their connected sum. So, the reef
and granny knots both have Alexander polynomial .
We consider three knots L+, L0, and L– that differ only at one crossing. The
knots L+ and L– differ in which part of the knot makes the over-crossing and
the knot L0 has no crossing at all at this point (Figure 46).
We will shortly use the skein relation to calculate the Jones polynomial of
the trefoil, but an important point has so far been ignored. Even if L+ and L–
are knots, L0 need not be. As we will see in the example of the trefoil,
eliminating a crossing might disconnect a knot and create what is called a
link, which is just a collection of knots. The skein relation, and Jones
polynomial generally, should be seen as relating to such links.
The Jones polynomial can sometimes differentiate between a knot K and its
mirror image K* as the identity holds; so, we need to
specify the trefoil being considered as right-handed. In our calculation we’ll
need to consider various knots and links (Figure 47).
47. Simple examples of links (a) Right-handed trefoil, (b) Twisted unknot, (c) Unlinked circles,
(d) Linked circles.
We will take L+ as the trefoil and focus on the bottom-right of the three
crossings. L- is then the unknot, perhaps easiest seen by unpoking the
bottom part of the knot with a Reidemeister move. And L0 has become
disconnected making two linked circles as in Figure 47(d). Remembering
that the skein relation states
where Vlinked(x) is the Jones polynomial of two linked circles. Using the
skein relation a second time, we take L+ as the linked circles, and consider
the rightmost crossing. Then L– is two unlinked circles (Figure 47(c)) and
L0 is the unknot. This time the skein relation states
Finally, we apply the skein relation to the twisted unknot (Figure 47(b)) to
find out Vunlinked(x). If we take L+ as the twisted unknot then L– is another
twisted version of the unknot and L0 is two unlinked circles. The skein
relation then states
And from the earliest history of topology, connections with physics would
be apparent. In the 19th century Gauss and Maxwell would both recognize
such in electromagnetism—for example, in the study of the work done by a
magnetic pole moving in the presence of a wire carrying current, with
Gauss’s answer being in terms of a linking number for the pole’s path and
the wire. Both Gauss and Maxwell would bemoan the lack of progress with
the study of topology or ‘geometry of position’ as Maxwell referred to it at
the time.
The 20th century would develop a yet richer connection between topology
and physics. In 1965, Roger Penrose would use topological ideas to
demonstrate how the gravitational collapse of a massive star would lead to a
space-time singularity occurring, such as a black hole. The use of topology
meant that Penrose was able to impose qualitative assumptions about the
mass distribution, compared with earlier assumptions about the symmetric
distribution of matter that had been considered physically questionable.
The interaction between physics and topology would also not be one way. A
problem that was ostensibly in physics would become of interest to pure
mathematicians if it could be rephrased into mathematical language
involving mathematical objects. This was particularly the case with Yang–
Mills theory, a physical theory seeking to provide a unified description for
electromagnetism and the weak force. In 1983 Simon Donaldson would use
ideas from Yang–Mills theory to prove astonishing results about the
topology of four-dimensional manifolds.
An oriented knot K, with n crossings c1, c2 … cn, divides the plane into
regions, , including the outside of the knot. Figure 48
shows a trefoil labelled in this manner.
1817 Bolzano defines continuity, proves the Bolzano–Weierstrass and intermediate value theorems
1857 Riemann’s paper Theory of Abelian Functions makes Riemann surfaces more widely known
1861 Möbius gives a first sketch proof of the classification theorem for closed surfaces
1874 Klein shows orientable closed surfaces are homeomorphic if and only if they have equal genus
1907 Max Dehn and Poul Heegaard give the first rigorous proof of the classification theorem
1914 Hausdorff defines topological spaces in the seminal Grundzüge der Mengenlehre
1925 Morse publishes his paper Relation between the Critical Points. . .
1932 Čech defines higher homotopy groups, generalizing the fundamental group
1950 Hamming distance introduced in the paper Error Detecting and Error Correcting Codes
1986 Donaldson and Freedman win Fields Medals for their work on four-dimensional manifolds
Online references
3Blue1Brown, Who Cares about Topology?<http://www.youtube.com/watch?
v=AmgkSdhK4K8&t=236s>
The Geometry Junkyard, 20 Proofs of Euler’s Formula
<http://www.ics.uci.edu/~eppstein/junkyard/euler/all.html>
Index
Note: For the benefit of digital users, indexed terms that span two pages (e.g., 52–53) may, on
occasion, appear on only one of those pages.
A
Alexander, James 114, 118, 123, 131
Alexander polynomial 123–124, 131
algebraic topology 90, 103–114
alternating knot 117, 120
ambient isotopy 115–116
Analysis Situs 105, 128
B
Betti numbers 112
blancmange function 62–63, 90
Bolzano, Bernard 53, 61
Bolzano–Weierstrass theorem 84
boundedness theorem 61, 85, 93
Brouwer’s fixed point theorem 109
C
Cauchy, Augustin-Louis 9
chirality 117, 123
classification theorem for knots 116
classification theorem for surfaces 38–40
closed sets 81
closed surfaces 29
combinatorial topology 113
compactness 63, 84–85
complex numbers 40–43
connected components 19, 86
connected sums 31–33, 119, 124
connectedness 63, 85–87
continuity 48, 57–62, 69–73, 76–77
convex sets 87
convergence 80
critical points 91–95
cut points 88
D
degree-genus formula 46
Descartes, René 10
differential topology 90–96
dimension 19, 21
domain of a function 51, 63
dual polyhedra 13
E
equivalent metrics 73–76
Euler number 25, 32–33, 37–40, 90, 95, 99, 112–113
Euler, Leonhard 5–6
Euler’s formula 5–10
F
Flatland 21–22, 82–83
football, see truncated icosahedron
fractals 21
Fréchet, Maurice 67, 78, 85
functions 50–51, 64
fundamental group 103–108
fundamental theorem of algebra 42
G
Gauss, Carl 42, 128
general topology 21, 78, 90
genus 33
graph theory 15–18
groups 101
H
hairy ball theorem 96–99
Hamming distance 68
Hausdorff, Felix 78, 85, 90
Heine–Borel theorem 85
historical timeline 135–136
homeomorphic 3–5
homomorphism 109
homotopic 104
I
index of a singularity 98
intermediate value theorem 60, 86
invariants 5, 85, 87–9, 100, 112–113
J
Jones polynomial 124
Jordan curve theorem 19
Jordan, Camille 19
K
Klein bottle 36–37
knot group 121–123
knots 22–23, 115
L
Lakatos, Imre 10
limit 80
links 125–6
M
manifolds 47, 99, 112–114
Maxwell, James Clerk 95, 116, 118, 128
metric spaces 67–76
Möbius strip 33–35
Morse theory 95
O
one-sided surfaces 33–39
open balls 73
open sets 76
orientable surfaces 35, 39
P
path-connectedness 86
physics 43, 128–129
Platonic solids 6, 10–13
Poincaré conjecture 114
Poincaré–Hopf theorem 99
points at infinity 43–45
prime knots 119–120
projective plane 38
R
Reidemeister moves 117–118
Riemann surfaces 45
Riemann, Bernhard 40, 46, 128
S
simply connected 5, 89, 107
singularity 96
skein relation 125
smooth functions 90–91
subdivisions 29–31
subspaces 82–84
surfaces 24–47
T
topological data analysis 129
topological spaces 78
topology (of open sets) 78
trefoil 22, 117
triangle inequality 67
truncated icosahedron 6, 14
U
Underground map 1–2
unknot 22, 116
unknotting problem 118
W
Weierstrass, Karl 53, 62, 128
Whitney, Hassler 47
Wirtinger presentation 121–123
ECONOMICS
A Very Short Introduction
Partha Dasgupta
Economics has the capacity to offer us deep insights into some of the most formidable
problems of life, and offer solutions to them too. Combining a global approach with
examples from everyday life, Partha Dasgupta describes the lives of two children who live
very different lives in different parts of the world: in the Mid-West USA and in Ethiopia. He
compares the obstacles facing them, and the processes that shape their lives, their families,
and their futures. He shows how economics uncovers these processes, finds explanations
for them, and how it forms policies and solutions.
www.oup.com/vsi
INFORMATION
A Very Short Introduction
Luciano Floridi
‘Splendidly pellucid.’
Steven Poole, The Guardian
www.oup.com/vsi
INNOVATION
A Very Short Introduction
Mark Dodgson & David Gann
This Very Short Introduction looks at what innovation is and why it affects us so profoundly.
It examines how it occurs, who stimulates it, how it is pursued, and what its outcomes are,
both positive and negative. Innovation is hugely challenging and failure is common, yet it is
essential to our social and economic progress. Mark Dodgson and David Gann consider the
extent to which our understanding of innovation developed over the past century and how it
might be used to interpret the global economy we all face in the future.
‘Innovation has always been fundamental to leadership, be it in the public or private arena.
This insightful book teaches lessons from the successes of the past, and spotlights the
challenges and the opportunities for innovation as we move from the industrial age to the
knowledge economy.’
Sanford, Senior Vice President, IBM
www.oup.com/vsi
NOTHING
A Very Short Introduction
Frank Close
What is ‘nothing’? What remains when you take all the matter away? Can empty space - a
void - exist? This Very Short Introduction explores the science and history of the elusive
void: from Aristotle’s theories to black holes and quantum particles, and why the latest
discoveries about the vacuum tell us extraordinary things about the cosmos. Frank Close
tells the story of how scientists have explored the elusive void, and the rich discoveries that
they have made there. He takes the reader on a lively and accessible history through
ancient ideas and cultural superstitions to the frontiers of current research.
‘An accessible and entertaining read for layperson and scientist alike.’
Physics World
www.oup.com/vsi
NUCLEAR POWER
A Very Short Introduction
Maxwell Irvine
The term ‘nuclear power’ causes anxiety in many people and there is confusion concerning
the nature and extent of the associated risks. Here, Maxwell Irvine presents a concise
introduction to the development of nuclear physics leading up to the emergence of the
nuclear power industry. He discusses the nature of nuclear energy and deals with various
aspects of public concern, considering the risks of nuclear safety, the cost of its
development, and waste disposal. Dispelling some of the widespread confusion about
nuclear energy, Irvine considers the relevance of nuclear power, the potential of nuclear
fusion, and encourages informed debate about its potential.
www.oup.com/vsi
NUMBERS
A Very Short Introduction
Peter M. Higgins
Numbers are integral to our everyday lives and feature in everything we do. In this Very
Short Introduction Peter M. Higgins, the renowned mathematics writer unravels the world of
numbers; demonstrating its richness, and providing a comprehensive view of the idea of the
number. Higgins paints a picture of the number world, considering how the modern number
system matured over centuries. Explaining the various number types and showing how they
behave, he introduces key concepts such as integers, fractions, real numbers, and
imaginary numbers. By approaching the topic in a non-technical way and emphasising the
basic principles and interactions of numbers with mathematics and science, Higgins also
demonstrates the practical interactions and modern applications, such as encryption of
confidential data on the internet.
www.oup.com/vsi
TOCQUEVILLE
A Very Short Introduction
Harvey Mansfield
No one has ever described American democracy with more accurate insight or more
profoundly than Alexis de Tocqueville. After meeting with Americans on extensive travels in
the United States, and intense study of documents and authorities, he authored the
landmark Democracy in America, publishing its two volumes in 1835 and 1840. Ever since,
this book has been the best source for every serious attempt to understand America and
democracy itself. Yet Tocqueville himself remains a mystery behind the elegance of his
style. In this Very Short Introduction, Harvey Mansfield addresses his subject as a thinker,
clearly and incisively exploring Tocqueville’s writings-not only his masterpiece, but also his
secret Recollections, intended for posterity alone, and his unfinished work on his native
France, The Old Regime and the Revolution.
www.oup.com/vsi