Topology As Fluid Geometry Two Dimensional Spaces Volume 2 PDF
Topology As Fluid Geometry Two Dimensional Spaces Volume 2 PDF
Topology As Fluid Geometry Two Dimensional Spaces Volume 2 PDF
TOPOLOGY AS FLUID
GEOMETRY
James W. Cannon
Two-Dimensional Spaces, Volume 2
TOPOLOGY AS FLUID
GEOMETRY
Two-Dimensional Spaces, Volume 2
TOPOLOGY AS FLUID
GEOMETRY
James W. Cannon
Copying and reprinting. Individual readers of this publication, and nonprofit libraries acting
for them, are permitted to make fair use of the material, such as to copy select pages for use
in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Permissions to reuse
portions of AMS publication content are handled by Copyright Clearance Center’s RightsLink
service. For more information, please visit: http://www.ams.org/rightslink.
Send requests for translation rights and licensed reprints to [email protected].
Excluded from these provisions is material for which the author holds copyright. In such cases,
requests for permission to reuse or reprint material should be addressed directly to the author(s).
Copyright ownership is indicated on the copyright page, or on the lower right-hand corner of the
first page of each article within proceedings volumes.
c 2017 by the American Mathematical Society. All rights reserved.
The American Mathematical Society retains all rights
except those granted to the United States Government.
Printed in the United States of America.
∞ The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at http://www.ams.org/
10 9 8 7 6 5 4 3 2 1 22 21 20 19 18 17
Contents
Chapter 3. Tools 29
3.1. Polyhedral complexes 29
3.2. Urysohn’s Lemma and the Tietze Extension Theorem 31
3.3. Set Convergence 34
3.4. Exercises 36
Chapter 6. The Arc, the Simple Closed Curve, and the Cantor Set 57
6.1. Characterizing the Arc and Simple Closed Curve 57
6.2. The Cantor Set and Its Characterization 61
6.3. Interesting Cantor Sets 63
v
vi CONTENTS
Chapter 9. 2-Manifolds 91
9.1. Definition and Examples 91
9.2. Exercises 91
Volume 2: The topology of the plane, with all of the standard theorems of
1 and 2-dimensional topology, the Fundamental Theorem of Algebra, the Brouwer
Fixed-Point Theorem, space-filling curves, curves of positive area, the Jordan Curve
Theorem, the topological characterization of the plane, the Schoenflies Theorem,
the R. L. Moore Decomposition Theorem, the Open Mapping Theorem, the trian-
gulation of 2-manifolds, the classification of 2-manifolds via orientation and Euler
characteristic, dimension theory.
Volume 3: An introduction to non-Euclidean geometry and curvature. What
is the analogy between the standard trigonometric functions and the hyperbolic
trig functions? Why is non-Euclidean geometry called hyperbolic? What are the
gross intuitive differences between Euclidean and hyperbolic geometry?
The approach to curvature is backwards to that of Gauss, with definitions
that are obviously invariant under bending, with the intent that curvature should
obviously measure the degree to which a surface cannot be flattened into the plane.
Gauss’s Theorema Egregium then comes at the end of the discussion.
Prerequisites: An undergraduate student with a reasonable memory of cal-
culus and linear algebra, but with no fear of proofs, should be able to understand
almost all of the first volume. A student with the rudiments of topology—open and
closed sets, continuous functions, compact sets and uniform continuity—should be
able to understand almost all of the second volume with the exeption of a little
bit of algebraic topology used to prove results that are intuitively reasonable and
can be assumed if necessary. The final volume should be well within the reach of
someone who is comfortable with integration and change of variables. We will make
an attempt in many places to review the tools needed.
Comments on exercises: Most exercises are interlaced with the text in those
places where the development suggests them. They are an essential part of the
text, and the reader should at least make note of their content. Exercise sections
which appear at the end of most chapters refer back to these exercises, sometimes
with hints, occasionally with solutions, and sometimes add additional exercises.
Readers should try as many exercises as attract them, first without looking at hints
or solutions.
Comments on difficulty: Typically, sections and chapters become more diffi-
cult toward the end. Don’t be afraid to quit a chapter when it becomes too difficult.
Digest as much as interests you and move on to the next chapter or section.
Comments on the bibliography: The book was written with very little
direct reference to sources, and many of the proofs may therefore differ from the
standard ones. But there are many wonderful books and wonderful teachers that
we can learn from. I have therefore collected an annotated bibliography that you
may want to explore. I particularly recommend [1, G. H. Hardy, A Mathematician’s
Apology], [2, G. Pólya, How to Solve It], and [3, T. W. Körner, The Pleasure of
Counting], just for fun, light reading. For a bit of hero worship, I also recommend
the biographical references [21, E. T. Bell, Men of Mathematics], [22, C. Henrion,
Women of Mathematics], and [23, W. Dunham, Journey Through Genius]. And
I have to thank my particular heroes: my brother Larry, who taught me about
uncountable sets, space-filling curves, and mathematical induction; Georg Pólya,
who invited me into his home and showed me his mathematical notebooks; my ad-
visor C. E. Burgess, who introduced me to the wonders of Texas-style mathematics;
R. H. Bing, whose Sling, Dogbone Space, Hooked Rug, Baseball Move, epslums and
PREFACE TO THE THREE VOLUME SET xi
deltas, and Crumpled Cubes added color and wonder to the study of topology; and
W. P. Thurston, who often made me feel like Gary Larson’s character of little brain
(“Stop, professor, my brain is full.”) They were all kind and encouraging to me.
And then there are those whom I only know from their writing: especially Euclid,
Archimedes, Gauss, Hilbert, and Poincaré.
Finally, I must thank Bill Floyd and Walter Parry for more than three decades
of mathematical fun. When we would get together, we would work hard every
morning, then talk mathematics for the rest of the day as we hiked the cities, coun-
trysides, mountains, and woods of Utah, Virginia, Michigan, Minnesota, England,
France, and any other place we could manage to get together. And special thanks
to Bill for cleaning up and improving almost all of those figures in these books
which he had not himself originally drawn.
Preface to Volume 2
The first of three volumes in this set was devoted to the measurement of lengths
and areas, and to some of the consequences that study had in number theory,
algebra, and analysis. Euclid was able to solve quadratic equations by geometric
construction. But when mathematicians tried to extend those results to equations
of higher degree and to differential equations, a number of fascinating difficulties
arose, all involving limits and continuity, best modelled by topology.
In this second volume we assume that the reader has had a first course in
topology and is comfortable with open and closed sets, connected sets, compact
sets, limits, and continuity. Two good references are W. S. Massey [25] and J. R.
Munkres [24].
The following discoveries led to the topics of this second volume.
(1) The solution of cubic and quartic equations required serious consideration of
complex numbers, thought at first to be mysterious. But the mystery disappeared
when it was seen that complex numbers simply model the Euclidean plane. Abel
and Galois proved that equations of degrees 5 and higher could not be solved in
the relatively simple manner by formula as had sufficed in equations of degrees 1
through 4. But Gauss, without giving explicit solutions, managed to prove the
Fundamental Theorem of Algebra that ensured that complex numbers sufficed for
their solution. Gauss gave proofs involving the geometry and topology of the plane.
(2) Newton showed that the study of motion could be greatly simplified if, in-
stead of examining standard equations, one examined differential equations. Prov-
ing the existence of solutions to rather general differential equations led to problems
in topology. One of the standard proof techniques involves Brouwer’s Fixed Point
Theorem. This volume proves that theorem in dimension 2 and outlines the proof
in general dimensions.
(3) Descartes demonstrated that mechanical devices other than straight edge
and compass can construct curves of very high degree. Once curves of very general
form are accepted as interesting, further delicate questions of length and area arise:
finite curves of infinite length, finite curves of positive area, space filling curves,
disks whose interiors have smaller areas than their closures, 0-dimensional sets
through which no light rays can penetrate, continuous functions that are nowhere
differentiable, sets of fractional dimension. This volume gives examples of many of
these phenomena.
(4) The study of solutions to equations became more unified when all variables
were considered to be complex variables. Riemann modelled complex curves by
surfaces, which are 2-dimensional manifolds and are called Riemann surfaces. The
analysis of 2-dimensional manifolds led naturally to notions, such as triangulation,
genus, and Euler characteristic. These notions are explained in this volume.
xiii
xiv PREFACE TO VOLUME 2
All of these considerations required the study of limits and continuity, and the
abstract notion that models limits and continuity in their most general settings is
the notion of topology. Henri Poincaré wrote:
As for me, all of the diverse paths which I have successively
followed have led me to topology. I have needed the gifts of this
science to pursue my studies of the curves defined by differential
equations and for the generalization to differential equations of
higher order, and, in particular, to those of the three body
problem. I have needed topology for the study of nonuniform
functions of two variables. I have needed it for the study of
the periods of multiple integrals and for the application of that
study to the expansion of perturbed functions. Finally, I have
glimpsed in topology a means to attack an important problem
in the theory of groups, the search for discrete or finite groups
contained in a given continuous group.
CHAPTER 1
Our first excursion into the topology of the plane will be in the proof of the
Fundamental Theorem of Algebra:
Theorem 1.1 (Fundamental Theorem of Algebra). If f (x) = xn + an−1 xn−1 +
· · · + a1 x + a0 = 0 is a polynomial equation in the unknown x and if the constant
coefficients an−1 , . . ., a1 , a0 are complex numbers, then there is a complex number
x = α that satisfies the equation: f (α) = 0. (Of course, since the real numbers are
a subset of the complex numbers (the line lies in the plane), these coefficients are
allowed to be real numbers.)
The Greeks solved linear and quadratic equations geometrically. They rejected
solutions that were not real numbers as being of no application or interest. Complex
numbers were first acknowledged as important in the solution of cubic equations.
General solution formulas for the cubic and quartic equations were found only by
great effort and cleverness.
Georg Pólya wrote to me when I was a young Mormon missionary in Austria.
He said that I should solve a hard mathematical problem every week so that I
wouldn’t rust (“Wer rastet, der rostet”). He also gave me a list of books that I
might find in a used bookstore. I learned the following argument from one of them:
Here is a general solution to the cubic equation, f (x) = x3 + ax2 + bx + c = 0:
We simplify the equation by translation, setting x = z + k. By subsititution, we
find
f (x) = z 3 + (3k + a)z 2 + (3k2 + 2ak + b)z + (k3 + ak2 + bk + c) = 0.
If we choose k = −a/3, the equation has the form
z 3 + pz + q = 0.
Setting z = u + v and substituting, we find
(u3 + v 3 + q) + (3uv + p)(u + v) = 0.
If we are able to choose u and v so that
u3 + v 3 + q = 0 and 3uv + p = 0,
the equation will be satisfied. We can solve this pair of equations for u and v in
the standard way by substitution:
u = −p/3v and, consequently,
3
−p
+ v 3 + q = 0.
3v
1
2 1. THE FUNDAMENTAL THEOREM OF ALGEBRA
Multiplying by v 3 ,
3
−p
v 6 + qv 3 + = 0.
3
Though this equation has degree 6 in the unknown v, it is only quadratic in the
unknown v 3 , so that by the quadratic formula,
3
3 2
q 1 p q q p
v =− ±
3
q2 + 4 =− ± + .
2 2 27 2 2 3
Assuming that we know how to take cube roots, we obtain v, from which we find
u3 = −v 3 − q or u = −p/3v, z = u + v, and x = z − (a/3).
For example, if√f (x) = x3 − 15x − 4 (already simplified), we have p = −15,
q = −4, v 3 = 2 + 4 − 125 = 2 + 11i, and u3 = 2 − 11i. But (2 + i)3 = 2 + 11i
and (2 − i)3 = 2 − 11i. The sum of these cube roots is (2 + i) + (2 − i) = 4,
which is a real root of the original equation. The fact that real roots could be
mediated by formulas involving complex numbers was a huge motivating factor for
the acceptance of complex numbers.
We have ignored some of the obvious difficulties: How do we take cube roots?
Further, if we take the three cube roots of v 3 and three of u3 , we would be able
to put together 9 possibilities for z, and there can only be three solutions to the
original equation. To choose those that are in fact roots, we must satisfy the more
restrictive equation, u = −p/3v. In the example,
−p 15 6 − 3i 15 · (6 − 3i)
= · = = 2 − i.
3(2 + i) 6 + 3i 6 − 3i 36 + 9
Mathematicians hoped to find similar solutions for polynomial equations of
higher degree, solutions that only required the arithmetic operations of addition,
subtraction, multiplication, and division together with the extraction of roots. This
hope was dashed by the work of N. H. Abel and E. Galois, who showed that the
four arithmetic operations and extraction of roots were inadequate in general for
expressing the roots of a quintic equation in terms of its coefficients.
I would love to include their proofs here, but Abel’s proof required twenty pages
of work and Galois’s development requires a substantial development of the theory
of fields and finite groups. Instead, we shall only prove the Fundamental Theorem
of Algebra. We first need to review some fundamentals from the arithmetic of
complex numbers.
(a + c) + (b + d)i
c + di
a + bi
(0, 0)
to think of the projection principle as determining the effect on lengths when one
length is projected orthogonally onto another.
The projection principle. Consider a right triangle with one of its acute
angles equal to α, and suppose that the length of the hypotenuse is r. Then the
leg adjacent to the angle α has length (cos α) · r, and the leg opposite the angle α
has length (sin α) · r.
To obtain the sum formulas for sin and cos, we apply the projection principle to
the following two diagrams; see Figure 4. Each term of the sum formulas represents
a well-defined geometric segment in the diagrams.
4 1. THE FUNDAMENTAL THEOREM OF ALGEBRA
(sin α) · r
(cos α) · r
α
s(α)s(β) s(α)s(β)
s(α)c(β)
s(α)c(β) s(α)
s(α)
1 1
c(α) c(α)
α c(α)s(β) α c(α)s(β)
β β
(0, 0) c(α)c(β) (0, 0) c(α)c(β)
We use c and s as shorthand for cos(α + β) and sin(α + β). We use c(α) and
s(α) as shorthand for cos(α) and sin(α), and similarly for cos(β) and sin(β). We
have drawn two right triangles in each of the diagrams, one with angle α, the other
with angle β. We have scaled the triangles so that the larger one has hypotenuse
1 so that the vertex emphasized by the large dot has coordinates (c, s). All of the
other entries are consequences of the projection principle. From the figures, it is
clear that
c = c(α)c(β) − s(α)s(α), or
cos(α + β) = cos(α) cos(β) − sin(α) sin(β), and similarly
sin(α + β) = cos(α) sin(β) + sin(α) cos(β).
These are the sum formulas for sin and cos.
Admittedly, the diagrams deal only with positive angles α and β whose sum is
≤ π, but all other angles can readily be reduced to these cases.
1.2. FIRST PROOF OF THE FUNDAMENTAL THEOREM 5
Sr × 0
Sr × 1
Since the set Sr × [0, 1] is compact and its image misses 0, there is a positive
distance from the image to the origin. The map Q is uniformly continuous, so that,
if the triangles are small, so are the images of those triangles. We may move the
images of the vertices a tiny bit so that the images of the vertices of each triangle are
vertices of a rectilinear triangle in the plane. We replace each (possibly curvilinear
and singular) image triangle by the corresponding rectilinear triangle spanned by
the new image vertices.
We take care in the adjustment that the following conditions are satisfied. The
vertex images from Sr × {0} are not moved. The image of the entire set Sr × {1}
is still very small. No adjusted triangle hits the origin 0.
1.3. SECOND PROOF 9
We now pick a ray R from 0 to ∞ in the plane in such a way that it misses all
of the new vertex images and misses the (very small) image of Sr × {1}. We lose
no generality in assuming that R is the positive x-axis.
We complete the proof by counting the number of times various triangle edges
intersect and cross the ray R. We orient the edges of Sr ×{0} in the counterclockwise
direction. The image of Sr ×{0} intersects R exactly n times, and each time crosses
R from the lower side to the upper side. We assign the crossing number +1 to each
oriented edge that crosses R from the bottom to the top. When we later come
upon oriented edges that cross R from top to bottom, we assign such crossings the
crossing number −1. See Figure 7. If an edge misses R, we assign it number 0.
The sum of crossing numbers for the image of Sr × {0} is n.
We delete triangles one at a time, layer by layer, as indicated by Figure 6. We
first remove the green triangles, then the orange triangles, etc.
Sr × 0
Sr × 1
Each time a triangle is deleted, we obtain a new boundary curve. Either one
oriented boundary edge is replaced by two new oriented boundary edges or two ori-
ented boundary edges are replaced by one new oriented boundary edge. It is easy
to see that the algebraic sum of assigned crossing numbers remains unchanged with
each triangle deletion since the sum obviously remains unchanged on each triangle.
In order to see this, examine Figure 7.
Eventually, one moves through the diagram from the image of Sr × {0}, which
has algebraic sum n, to the image of Sr × {1}, which has algebraic sum 0, a con-
tradiction since the algebraic sum remains unchanged.
With this contradiction, the proof of the theorem is complete.
1.4. Exercises
Abel and Galois proved that there are no formulas like the quadratic formula
for finding the roots of polynomials with real or complex coefficients of degrees
≥ 5. The Fundamental Theorem of Algebra, for which we have just given two
proofs, shows that there are always roots, in fact, if we count possible multiplicity
of roots, exactly as many roots as the degree of the polynomial. The problem
becomes one of finding the roots, or at least approximating the roots to a desired
degree of accuracy. There is a large literature concerning efficient ways of finding
the roots, some of which have been programmed into most pocket calculators. In
the following exercises, we suggest exploring one of the standard methods, namely,
Newton’s method.
1.1. (The equation of the tangent line) Let p(x) be a polynomial with real
coefficients and let xn be some real number. Find the equation of the tangent line
to the graph of p(x) at the point (xn , p(xn )).
Answer: y = p(xn ) + p (xn ) · (x − xn ).
1.2. Find the point xn+1 at which the tangent line crosses the x-axis.
Answer: xn+1 = xn − p(xn )/p (xn ).
Newton’s method: Newton notes that if xn is a root of p(x), then xn+1 = xn .
That is, a root of p(x) is a fixed point of the operation xn
→ xn+1 , and when xn is
very close to a root r of p(x), then xn+1 is much√closer to r.
1.3. Use Newton’s method to approximate 5. (That is, find an approximate
root to the equation x2 − 5 = 0.)
1.4. Apply Newton’s method in an attempt to find a root of the equation
x2 − 2x + 10 = 0. Why does the method fail?
1.5. Use the quadratic formula to find the two roots of x2 − 2x + 10.
Newton’s method often works even when there are no real roots. Use the same
iterative formula xn+1 = xn − p(xn )/p (xn ), but start the process with x0 equal to
some nonreal complex number. This requires the ability to do complex arithmetic.
1.6. Apply Newton’s method to the polynomial x2 − 2x + 10 but with x0 equal
to the complex number i.
1.7. Construct a computer program to color the plane. Starting with an initial
complex number x0 representing a pixel in the plane, iterate Newton’s process
enough times that the result seems to be getting close to one of the roots of the
polynomial and color the original pixel a color assigned to that root. If the method
doesn’t seem to converge given that initial value, leave that pixel uncolored. Iterate
the process with different initial pixels.
CHAPTER 2
As the climax of this chapter, we will prove the Brouwer Fixed Point Theorem.
Our proof will be complete only in dimension 2, but the technique we use is valid in
all dimensions and lacks only a tiny bit of technique to complete, technique that we
will outline in exercises. This theorem is one of the two basic tools used in the proof
of existence of solutions to differential equations, the other being the contraction
mapping principle. A generalized version of the theorem is used in game theory to
prove the existence of optima.
In order to completely understand this chapter, the reader should have a rea-
sonable grasp of limits, continuity, open sets, and closed sets in Euclidean space.
night. The next morning he left the mountain top, again at 6:00 AM and returned
down the trail to his monastery at the foot of the mountain. Explain why there was
a time of day and a place on the trail so that the monk was at that place on the
trail at that same time on each of the two days. [The answer to the puzzle given
in the story was “The monk had to meet himself.”]
The proof of the Brouwer Fixed Point Theorem, as well as the proof of many
theorems in topology, relies on the introduction of extra structure to the problem.
We will illustrate the value of additional structure in a series of arguments. The
first will be an elementary puzzle involving the checkerboard. The second will use a
childhood puzzle to motivate the use of the sign of a permutation; this notion will be
important when we treat the mathematical notion of right and left, or orientation
in classifying 2-dimensional surfaces. The third will show how to justify the (?)
obvious (?) fact that a polygonal closed curve in the plane separates the plane
into two pieces, an inside and an outside; this argument will introduce the powerful
topological tool of general position. The idea of general position, together with a
trick called the one-ended arc trick, will be used to prove the famous No Retraction
Theorem. Finally, we will easily deduce the Brouwer Fixed Point Theorem from
the No Retraction Theorem.
The puzzle consists of fifteen tabbed and slotted plastic tiles set in a plastic
case, with one blank slot (pictured as black) into which adjacent tiles can be slid.
See Figures 3 and 4.
2.3. TWO ELEMENTARY PROBLEMS 13
The goal was to transform a scrambled arrangement of tiles into the unscram-
bled version. This puzzle was, for a time, the rage — just as the Rubik’s Cube had
its period — and the puzzle is still manufactured and sold as a children’s diversion.
I played with the puzzle at church when I was a child when I was bored with the
sermon. Puzzle enthusiasts began to ask whether all arrangements of tiles could
be realized. In particular, they tried unsuccessfully to transform the standard ar-
rangement into the reverse arrangement. See Figure 5.
Books were written about the puzzle. A museum at the Indiana University,
Bloomington, collects and displays the wonderful artifacts concerning this puzzle.
Why, or why not, can the reverse position be realized?
1 2 3 4
5 6 7 8
9 10 11 12
13 14 15
1 2 3 4
5 6 8
9 10 7 11
13 14 15 12
Second approximation for the tiling problem. Introduce extra structure to both
the board and to the domino by coloring every other square black. See Figures 6
and 7.
Solution to the tiling problem: Every domino, whether horizontally placed or
vertically placed, will cover one white and one black square. Thus, if the tiling
exists, there must be the same number of black and white squares. But there are
32 black squares and only 30 white squares. Therefore the tiling is impossible.
15 14 13 12
11 10 9 8
7 6 5 4
3 2 1
There are standard ways of picturing a permutation p. For example, the 2-row
form
1 2 3 4 5 6 7 8
p=
7 2 4 3 1 8 5 6
exhibits each i above its image p(i), so that, for example, p(1) = 7 and p(7) = 5.
The disjoint cycle form
(1 7 5)(2)(3 4)(6 8)
indicates that p naturally divides the elements of Sn into circles or cycles:
p : 1
→ 7
→ 5
→ 1,
p : 2
→ 2,
p : 3
→ 4
→ 3,
p : 6
→ 8
→ 6.
The cycle (1 7 5) is called a 3-cycle; the cycle (2) is a 1-cycle, and the cycles (3 4)
and (6 8) are 2-cycles. The 1-cycles of a permutation are often omitted in the
disjoint-cycle form. A 2-cycle is also called a transposition.
Each cycle can itself be considered as a permutation (a function), where all
elements of Sn not explicitly mentioned are to be considered as (omitted) 1-cycles.
Since functions are typically composed from right to left, cycles (as functions) can
also be composed, whether they are disjoint or not, to form new permutations. For
example,
(1 3 5 7)(2 5 3) = (2 7 1 3)(5).
Theorem 2.7. Every permutation of Sn can be realized as a product of 2-cycles
(transpositions).
Proof. It suffices to show that a cycle (a1 a2 . . . ak ) is a product of 2-cycles,
but
(a1 a2 . . . ak ) = (a1 ak )(a1 ak−1 ) · · · (a1 a3 )(a1 a2 ).
Definition 2.8. A permutation is even if it can be expressed as the product
of an even number of 2-cycles. A permutation is odd if it can be expressed as the
product of an odd number of 2-cycles.
Definition 2.9. Suppose that p is a permutation of Sn . Consider the un-
ordered pairs (i, j). We say that (i, j) is inverted by p if p changes the order of
i and j; that is, (i, j) is inverted by p if either (i < j and p(j) < p(i)) or (j < i
and p(i) < p(j)). Let inv(p) denote the number of unordered pairs (i, j) that are
inverted by p.
Here is an example, namely the permutation
1 2 3 4 5 6 7
.
5 1 6 3 7 4 2
If we draw arrows from each integer to its image spot, then the number of inversions
is the number of crossings of these arrows. See Figure 8.
Theorem 2.10. The permutation p is even iff inv(p) is even. The permutation
p is odd iff inv(p) is odd. (Consequently, no permutation is both even and odd.)
2.3. TWO ELEMENTARY PROBLEMS 17
1 2 3 4 5 6 7
5 1 6 3 7 4 2
Proof. It suffices to show that each 2-cycle changes the number of inversions
by an odd number. We start with the 2-row representation
1 ... i (i + 1) . . . (k − 1) k ... n
p(1) . . . p(i) p(i + 1) . . . p(k − 1) p(k) . . . p(n)
and compose with the transposition (p(i) p(k)), with i < k. The result, after the
composition is
1 ... i (i + 1) . . . (k − 1) k ... n
.
p(1) . . . p(k) p(i + 1) . . . p(k − 1) p(i) . . . p(n)
If k = i + 1, so that ai and ak are adjacent before the move, then the only
pair whose status of inversion is affected by the transposition is the pair (i, k): if
this pair is inverted before the transposition, then it is in order after; if it is in
order before the transposition, then it is inverted after. Thus the parity of inv(p)
is changed by this adjacency transposition. If there are = k − i − 1 elements
between ai and ak , then adjacency moves can move ai so that it is in position
k − 1 adjacent to ak , with changes in parity. Then ai and ak can be interchanged,
with a change of 1 in parity. Finally, ak can be moved back to position i by another
adjacency switches, with changes in parity. The result is 2 + 1 parity changes,
an odd number of changes. Hence, the original transposition resulted in a change
in the parity of inv(p).
We are now in the position to show that the reverse position cannot be reached
from the initial position.
We think of the blank slot as a tile labelled 16. We can view each position of
the puzzle as a permutation of the numbers 1 through 16. Each move in the puzzle
is a transposition that interchanges 16 with another tile. Hence, each move changes
the parity of the permutation. The tile 16 begins in a position in the checkerboard
pattern that is dark. Each move takes 16 from a dark position to a light position, or
vice versa. Hence, the position of 16 in the checkerboard pattern indicates whether
the permutation is even or odd: dark position = even permutation; light position
= odd permutation.
In the reverse position, tile 16 is in dark position (Figure 9); hence the permu-
tation, if attainable, must be even. We count the number of inversions. (Compare
with the original position, Figure 10. The number 15 is inverted with fourteen tiles,
the number 14 with 13 tiles, the number 13 with 12 tiles, etc. That is, the number
18 2. THE BROUWER FIXED POINT THEOREM
15 14 13 12
11 10 9 8
7 6 5 4
3 2 1
1 2 3 4
5 6 7 8
9 10 11 12
13 14 15
of inversions is
14 + 13 + 12 + 11 + 10 + 9 + 8 + 7 + 6 + 5 + 4 + 3 + 2 + 1,
which is odd. Thus this position is not attainable.
Exercise. Show that a tile position can be attained iff the parity of the permu-
tation agrees with the parity dictated by the position of tile 16.
we denote by S1 , can be placed in the plane. For the remainder of this section,
it is necessary to know about continuous functions.
2.4.1. Polygonal Simple Closed Curves in the Plane.
Definition 2.11. A simple closed curve in the plane is the image of a contin-
uous function J : S1 → R2 from the circle S1 into the plane R2 , where the function
J is one to one (that is, has no self-intersections).
For the next few theorems, we shall assume that the simple closed curve is
polygonal, that is, formed by a finite sequence of straight line segments. Figure 11
gives two examples, the first very simple, the second rather elaborate.
u
v
Example 1
Example 2
Figure 11. The simple and the complex simple closed curve
For Example 1, the triangle, there is evidently both an inside and an outside
to the curve. This is likewise true for a round circle. It is intuitively obvious that
every convoluted “circle” (simple closed curve) in the plane should have an inside
and an outside, but the proof that it is so becomes complex as we observe Example
2 and ask ourselves whether the points u and v that we have marked are inside the
curve or outside the curve.
Exercise 2.12. Is the point v inside the curve? The point u?
For a polygonal simple closed curve, there are at most two pieces in the com-
plement R2 \ J of J, for, locally, there are only two sides (Figure 12), and, as one
traverses the curve, the local sides near one point are connected to local sides of
surrounding points.
The question is only whether, when we follow those two local sides around
the full length of J, does the black Side 1 come back to connect again with the
Side 1, or does the black Side 1 come back to connect with grey Side 2 to form a
twisted Möbius band. We will give a proof of the following theorem. Although the
theorem seems obvious, the proof is not trivial and the techniques of its proof will
be used to prove the very important No Retraction Theorem and Brouwer Fixed
Point Theorem.
20 2. THE BROUWER FIXED POINT THEOREM
Side 1
Side 2 J
Figure 12. The two local sides of a polygonal simple closed curve
Theorem 2.13. Each polygonal simple closed curve J in the plane R2 separates
the plane. That is, the complement R2 \ J of J in R2 is not connected. In fact, it
has exactly two pieces (called components), an inside (called the interior) and the
outside (called the exterior).
Remark. The theorem is completely obvious for the triangle, and our proof
will make use of that fact. The proof that we will give is elementary in the sense
that it uses only basic geometric ideas, but the logic involved invokes ideas that,
in fully developed form, are very important in the development of geometric and
algebraic topology. We will try to point out the critical ideas as we go along. End
remark.
Here are a few of the basic ideas:
Idea (1) A triangle T in the plane has both an inside and an outside. Every
time you cross the boundary of the triangle you pass from outside to inside, or from
inside to outside. See Figure 13. If we traverse a polygonal simple closed curve
J that crosses ∂T at every intersection point, then every crossing point where J
crosses into T is followed by a subpath of J in J ∩ T that is terminated by a point
where J crosses the boundary ∂T out of T .
Proof (theorem). The proof is called the one-ended arc trick. We will show
that, unless the theorem is true, there is a finite edge path that has only one
endpoint, in contradiction to Idea (3) above. We therefore begin the proof by
assuming the ridiculous fact that the curve J does not separate the plane.
The proof requires that we introduce extra structure in our picture. This struc-
ture will take place in two different copies of the plane. We call the plane that
contains our polygonal simple closed curve J the image plane. The other plane
that we introduce we will call the model plane.
If our curve J contains n vertices, then we shall first construct n model triangles
in the model plane and then n corresponding image triangles in the image plane.
See Figures 14 and 15. After we have constructed the image triangles, we will
construct a second polygonal simple closed curve K in the image plane that is
in general position with respect to each of the n image triangles. See Figure 16.
Corresponding to each intersection arc of K with an image triangle, there is a
corresponding arc in the model triangle. It is in this collection of model arcs that
we will discover a one-ended arc, an obvious contradiction. We will conclude that
J must separate the image plane. (I find such an argument completely wild! I love
it.)
We first introduce the triangles in the model plane. If our first polygonal simple
closed curve has n vertices, then we pick n vertices on the unit circle S1 in the model
plane and join them by straight arcs to form a polygonal simple closed curve. We
then add a radius of the circle from each vertex A to the center V of the circle. We
obtain a model array of triangles ABV inside the disk of unit radius. It is in this
model that we will trace certain important paths:
S1
B
V
A
We map this model disk of triangles into the plane in the following way. We
map the outer boundary of the model to our original polygonal simple closed curve,
vertex to vertex, edge to edge. We map the center V of our model to some rather
arbitrary point V in the plane, subject only to the condition that it not be in any
of the lines defined by the edges of J. Then, given any edge e of J with vertices A
and B , the points A , B and V are the vertices of a triangle A B V in the plane.
The corresponding points A, B, and V of our model bound a triangle ABV in our
model. We map the image triangle A B V to the model triangle ABV linearly.
B
A
Z
V
While the model triangles intersect only along edges, the image triangles do
intersect along the corresponding common edge but may also fold back over each
other along that edge. For example, the triangles A B V and Z A V fold over each
other in the figure.
We have now completed the introduction of triangles into our problem. We
have still one additional structure to add to our picture. This is the addition of a
second polygonal simple closed curve in the image plane.
Since we are assuming that J does not separate the image plane, it does not
separate points just opposite each other across the arc A B . We may thus construct
a polygonal simple closed curve K in the image plane that intersects J only at one
point X ∈ A B where it crosses J.
This construction completes the added structure to our problem: n triangles
in the model plane, n triangles in the image plane, and a polygonal simple closed
curve K in the image plane.
2.4.2. The One-sided Arc Trick. We now examine the way K intersects
each of the image triangles. After a slight translation, we may assume that K misses
all of the image vertices and crosses each triangle edge at each point of intersection
(general position, Idea (2)).
Thus by Idea (1), if K intersects a triangle, it does so in a finite collection
of polygonal arcs, each with its endpoints on the boundary of the triangle. The
corresponding model triangle likewise is crossed by a corresponding finite collection
of polygonal arcs (Figure 17), pushed over from the image triangle by the linear
correspondence between the two triangles.
2.4. THREE ADVANCED PROBLEMS 23
B
X
A
Z
Image Model
B
X
A
V
The important property of these crossing arcs is the following. Whether adja-
cent triangles in the image plane fold back on each other or not, any intersection
arc from the one triangle that meets the common edge extends into the neighboring
24 2. THE BROUWER FIXED POINT THEOREM
triangle at that point. If the triangles did not fold on each other, then the arc is
simply an extension into the adjacent triangle. If the triangles did fold on each
other, then the arc enters the common edge in one triangle and folds right back on
top of itself in the adjacent triangle.
As a conseqence, in the model plane the intersection arcs form nonsingular
curves, which can only be finite edge paths or polygonal simple closed curves.
The key observation is this: Look at the intersection path in the model that
begins at the point X. Because X is an endpoint of that path, the path must have
a second endpoint by Idea (3). That endpoint must lie on a triangle edge and that
edge cannot be an interior edge of the model because intersection paths continue at
every interior edge. But no other boundary point is available since X is the only
point at which K hits a boundary edge.
We have discovered a one-ended arc, a contradiction.
Exercise 2.14. Prove that there is no polyhedral Möbius strip in the Euclidean
plane.
Theorem 2.16 (Brouwer Fixed Point Theorem). If D is the unit square disk,
and f : D → D is a continuous function that maps D into itself, then there is at
least one point x ∈ D such that f (x) = x.
2.4. THREE ADVANCED PROBLEMS 25
We will deduce the 2-dimensional Brouwer Fixed Point Theorem from the 2-
dimensional version of an equally famous theorem called the No Retraction Theo-
rem. We will prove the No Retraction Theorem by the one-ended arc trick.
Theorem 2.17 (No Retraction Theorem). If Bn is the unit n-dimensional ball
(the ball of radius 1), then it is impossible to find a continuous function f : Bn →
(Sn−1 = ∂Bn ) such that, for each x ∈ Sn−1 , f (x) = x. Such a map, if it existed,
would be called a retraction.
Proof of the No Retraction Theorem in dimension 2. In place of the
round disk B2 we use the square disk D to which it is topologically equivalent. As in
our previous proof using the one-ended arc trick, we make the ridiculous assumption
that the theorem is false. Hence there is a continuous function f : D → ∂D such
that, for each x ∈ ∂D, f (x) = x.
Again, we consider two planes: the model plane and the image plane. In the
symbol f : D → ∂D, we view the first D to be in the model plane and the second
D to be in the image plane. As before, we add structure to the model disk D by
subdividing it into very tiny triangles. See Figure 18.
We now use the continuous function f to carry the triangles of the square in
the model plane over to tiny triangles in the image plane. If T = ABC is any of the
tiny triangles, with vertices A, B, and C, then the images f (A), f (B), and f (C)
will be three points of ∂D that are very close to one another. We may move them
a tiny bit so that they are still near one another and still near ∂D, yet they are the
vertices of an actual triangle in the image plane. This process does not require that
we move any vertices that were originally in ∂D. In this manner, we modify all
26 2. THE BROUWER FIXED POINT THEOREM
those image vertices that must be moved. We may replace the original map f with
a new continuous function g : D → R2 that takes each triangle of the model space
linearly to the corresponding triangle of the image space. The new map perhaps
does not take D into ∂D, but it certainly takes D into the complement of the center
of the disk and does not move any point of ∂D.
We now have to add to the structure a path R analogous to the polygonal
simple closed curve K in the separation theorem. For R we take an infinite ray
that begins at the center of D in the image plane and misses every image vertex
under the map g, and passes onward to infinity. Then R either misses an image
triangle g(T ) entirely or else intersects g(T ) precisely in an arc that enters g(T )
through one edge and exits g(T ) through another edge. See Figure 19.
g(B)
g(A)
g(C)
The intersection arc can be carried back to an intersection arc in the model
triangle T = ABC by the linear correspondence between T and g(T ).
As in the previous proof, we concentrate on the union of the intersection arcs
as viewed in the model disk. As before, this union must be a finite collection of
finite edge paths, each having two endpoints, and a finite collection of polygonal
simple closed curves. As before, we consider that component that contains the
only intersection point of R with ∂D. That component must be an arc with two
endpoints by Idea (3). The second endpoint can only lie on the boundary of D, and
there is no other intersection with ∂D, a contradiction. That is, this component is
a one-ended arc.
As before, we conclude that the theorem is true, that our original assumption
was truly ridiculous.
Proof of the Brouwer Fixed Point Theorem. Since the round disk E
and the square disk D are topologically equivalent, we may replace D by E in both
the No Retraction Theorem and the Brouwer Fixed Point Theorem. We assume
2.5. EXERCISES 27
that, contrary to the latter, there is a continuous function g : E → E such that, for
each x ∈ E, g(x) = x. We then define a function f : E → ∂E as follows.
Consider the ray R(x) that begins at g(x) in E and passes through x on its
way to ∞. Define f (x) to be the last point of E in the ray R(x). Then the function
f : E → ∂E is a continuous function that fixes each point of ∂E, in contradiction
to the No Retraction Theorem.
2.5. Exercises
2.1. Solve Exercise 2.3 on page 11.
2.2. Solve Exercise 2.14 on page 24.
2.3. Solve Exercise 2.15 on page 24.
2.4. Generalize the No Retraction Theorem to dimension 3. Consider the unit
cube. Show how to divide the unit cube into tiny tetrahedra. View these tetrahedra
as lying in the model cube. Show how to carry those tetrahedra to tiny tetrahedra
in the image cube. Determine what the appropriate general position property is for
the intersection of a ray with a tetrahedron. Pull the intersection segments back
into the model cube. Show that you obtain a one-ended arc.
2.5. Contemplate what would have to be done to prove a No Retraction Theo-
rem in every dimension. Think how you would deduce a generalized Brouwer Fixed
Point Theorem in that dimension.
The Brouwer Fixed Point Theorem does not suggest a method for finding a
fixed point. If the map f : B2 → B2 is a contraction mapping, a particularly nice
property that we will now explain, then it is an easy matter to approximate a fixed
point. (Contraction mappings defined on other suitable spaces are one of the main
tools used to find solutions to differential equations.)
Definition 2.18. A map f : B2 → B2 is a contraction mapping if there is
a number 0 < λ < 1 such that, for each two points x, y ∈ B2 , the distances
d(f (x), f (y)) from f (x) to f (y) and d(x, y) from x to y satisfy the inequality
d(f (x), f (y)) ≤ λ · d(x, y).
In the following exercise you will need to use one of the fundamental properties
of a compact metric space: Every Cauchy sequence converges. (Recall that a
sequence x1 , x2 , . . . is a Cauchy sequence if, for each > 0, there is an integer N
such that n, m ≥ N implies d(xn , xm ) < .)
2.6. Prove that, if f : B2 → B2 is a contraction mapping, then f has a unique
fixed point x0 . If x1 ∈ B2 and if xn+1 = f (xn ) for each n ≥ 1, then the sequence
x1 , x2 , . . . is a Cauchy sequence converging to x0 .
2.7. Show that a tile position can be attained iff the parity of the permutation
agrees with the parity dictated by the position of tile 16.
2.8. Consider puzzles like the 15 puzzle, but with dimensions m × n. Which
configurations can be realized by such a puzzle?
2.9. Suppose that J is a polygonal simple closed curve in the plane bounding
a disk D. Show that D can be divided into triangles using only the vertices of J as
vertices of the triangles used.
2.10. By induction on the number of triangles used in the previous exercise,
show that D is homeomorphic to a single triangle T . (That is, you must prove the
existence of a continuous bijection f : D → T .)
CHAPTER 3
Tools
We take time to introduce three tools that are interesting in and of themselves.
The most useful additional structure added to problems in geometric topology
is the polyhedral complex; complexes are essentially used as finitely describable
approximations to spaces that have continuous structure.
The space being approximated can often be mapped into the finite polyhedral-
complex approximation by means of either Urysohn’s Lemma or the Tietze Exten-
sion Theorem.
Sequences of simple spaces often have interesting limit spaces. The limiting
process is described by Set Convergence.
A + λ · (B − A) = (1 − λ) · A + λ · B
Exercise 3.4. If a, b, c are given vectors, then the convex hull of the set {a, b, c}
consists of the sums
λ1 a + λ2 b + λ3 c,
where each λi ≥ 0 and λ1 + λ2 + λ3 = 1.
Exercise 3.5. The smallest 2-dimensional plane in 3-dimensional space con-
taining (noncollinear) vectors a, b, c consists of the sums
λ1 a + λ2 b + λ3 c,
where λ1 + λ2 + λ3 = 1.
The simplest convex sets are the simplexes. A 1-simplex is a closed interval. A
2-simplex is a triangle (together with its interior). A 3-simplex is a solid tetrahe-
dron.
Definition 3.6. Let O denote the origin in Rn . Let v1 , v2 , . . ., vn denote n
linearly independent vectors. Then the convex hull
σ = H({O, v1 , v2 , . . . , vn }) = {λi vi | li ≥ 0 and λi ≤ 1}
i>0 i>0
is called an n-simplex.
If w0 , w1 , w2 , . . ., wn are any n + 1 vectors, then we may map σ linearly onto
the convex hull H = H({w0 , w1 , w2 , . . . , wn }) by the formula
f (λ1 v1 + λ2 v2 + · · · + λn vn ) = λ0 w0 + λ1 w1 + · · · + λn wn ,
3.2. URYSOHN’S LEMMA AND THE TIETZE EXTENSION THEOREM 31
where λ0 = 1 − i>0 λi . If the map f is 1 to 1, so that it is a homeomorphism,
then f (σ) is also called an n-simplex.
The vectors w0 , w1 , . . ., wn are called the vertices of the simplex. The faces of
the simplex are the convex hulls of the subsets of the vertices. Consequently, the
endpoints of a 1-simplex are faces, as are the edges and vertices of a 2-simplex, and
the 4 triangular faces, the 6 edges, and the 4 vertices of a 3-simplex, and, of course,
the simplex σ is a face of itself.
Definition 3.7. A simplicial complex K is a collection of simplexes such that,
if two of the simplexes intersect, then they do so in a face of each. The union
|K| of the simplexes of K is called the carrier of K. We assign a topology to |K|
by declaring a subset to be closed (open) if its intersection with each face of each
simplex of K is closed (open) in that face.
Definition 3.8. We say that a space X is a polyhedron if it is the carrier of a
simplicial complex. We say X is triangulable if it is homeomorphic to a polyhedron.
A topological triangulation of X is a homeomorphism h : |K| → X from the carrier
of some simplicial complex K.
Proceding inductively, we divide each of the intervals I(n, k) into two subin-
tervals I = I(n + 1, 2k − 1) and J = I(n + 1, 2k) of equal length. We set A(I) =
f −1 (I) ∪ (X(n, k) ∩ X(n, k − 1)) and A(J) = f −1 (J) ∪ (X(n, k) ∩ X(n, k + 1)). We
define X(I) = {x ∈ X(n, k) | d(x, A(I)) ≤ d(x, A(J))} and X(J) = {x ∈ X(n, k) |
d(x, A(J)) ≤ d(x, A(I))}. We set X(n+1, 2k −1) = X(I) and X(n+1, 2k) = X(J).
Finally, every point x ∈ X is in a descending sequence of closed sets X(1, i1 ) ⊃
X(2, i2 ) ⊃ X(3, i3 ) ⊃ · · · , and the corresponding intervals I(1, i1 ) ⊃ I(2, i2 ) ⊃
I(3, i3 ) ⊃ · · · intersect at a single point y. We define F (x) = y. It is an easy
matter to see that F is continuous and is an extension of the original map f .
The general proof of both Urysohn’s Lemma and Tietze’s Extension Theorem
involves the construction of sets like the X(i, k)’s of the metric proof of Tietze’s
Extension Theorem. For Urysohn’s Lemma, the construction can be quite loose
and easy. For Tietze’s theorem, the construction requires more care.
Proof of Urysohn’s Lemma in the general case. Since X is normal
and since A and B are disjoint closed sets, there are disjoint open sets U ⊃ A
and V ⊃ B. Let X(1, 1) = X \ V and X(1, 2) = X \ U .
Assume inductively that closed sets X(n, 1), X(n, 2), . . ., X(n, 2n ) have been
chosen whose union is X. Assume further that X(n, j) and X(n, k) are disjoint
unless |j − k| ≤ 1. Then, for each k, the sets X(n, 1) ∪ · · · ∪ X(n, k − 1) and
X(n, k +1)∪· · ·∪X(n, 2n ) are disjoint closed sets. By normality, there exist disjoint
open sets U containing X(n, 1) ∪ · · · ∪ X(n, k − 1) and V containing X(n, k + 1) ∪
· · · ∪ X(n, 2n ). Let X(n, 2k − 1) = X(n, k) \ V and X(n, 2k) = X(n, k) \ U .
As in the proof of the metric Tietze theorem, every point x ∈ X is in a de-
scending sequence of closed sets X(1, i1 ) ⊃ X(2, i2 ) ⊃ X(3, i3 ) ⊃ · · · , and the
corresponding intervals I(1, i1 ) ⊃ I(2, i2 ) ⊃ I(3, i3 ⊃ · · · intersect at a single point
y. We define F (x) = y. It is an easy matter to see that F is continuous and is an
extension of the original map f .
In general, the set U (i) must contain the closure of U (i − 1) and f −1 ([a, xi ])
and (using normality) have closure missing both the closure of V (i − 1) and the set
f −1 (J).
Likewise, the set V (j) must contain the closure of V (j − 1) and f −1 ([yj , c]) and
(using normality) have closure missing both the closure of U (j) and the set f −1 (I).
The sets U and V so defined satisfy the requirements of the theorem. This
construction completes the proof of the Tietze Extension Theorem.
disjoint open sets U and V containing A and B, respectively. Since X = lim inf Xn ,
we may assume that every one of the sets Xn intersects both U and V . Since Xn
is connected, it must contain a point xn of S \ (U ∪ V ). Passing to a subsequence
if necessary, we may assume the sequence xn converges to a point x ∈ / U ∪ V . This
point must be in lim sup Xn = X, a contradiction.
Remark. We need not assume that the sets Xn are actually connected. It
suffices to assume that they are more and more nearly connected as n → ∞. Below
is the appropriate defintion. End remark.
Definition 3.18. A set Y is said to be δ-connected, if each two points x and
y of Y are connected by a chain x = x0 , x1 , . . . , xn = y of points of Y having the
property that the distance between xi and xi−1 is < δ for each i = 1, . . . , n.
Theorem 3.19. If the space S is compact, the sequence {X1 , X2 , . . .} of non-
empty sets converges to a set X, and, for each n, the set Xn is 1/n-connected, then
the limit X is nonempty, closed, and connected.
Proof. The limit is closed, hence compact. If X is not connected, say X =
A ∪ B, with A and B separated, then A and B lie in disjoint open sets U and V
that have disjoint closures. There is a positive distance δ between the sets U and
V . When 1/n < δ, any 1/n chain joining points of U and V must have a point xn in
the complement of U ∪ V . A subsequence converges to a point x ∈ (lim sup Xn ) \ X,
a contradiction.
Since the compact subsets of a separable metric space can be metrized (the
Hausdorff Metric) so as to form a separable metric space, the convergence theorem
has interesting applications to that space.
Definition 3.20 (Definition of the Hausdorff Metric). If X is a separable
metric space and A and B are compact subsets of X, then we define
d(A, B) = max sup{d(a, B) | a ∈ A}, sup{d(b, A) | b ∈ B} .
3.4. Exercises
3.1. Solve Exercise 3.3 on page 29.
3.2. Solve Exercise 3.4 on page 30.
3.3. Solve Exercise 3.5 on page 30.
3.4. Solve Exercise 3.1 on page 29.
3.5. Solve Exercise 3.9 on page 31.
3.6. Solve Exercise 3.21 on page 35.
3.7. Solve Exercise 3.22 on page 35.
3.8. Solve Exercise 3.23 on page 35.
[Hint: Consider the function f : X → Y as the compact subset {(x, f (x)) | x ∈
X} of the separable metric space X × Y .]
3.9. Solve Exercise 3.24 on page 35.
3.10. Solve Exercise 3.25 on page 35.
The set convergence theorem has some rather surprising consequences. Certain
spaces of functions can be realized as separable metric spaces:
3.11. If F = {fα : S1 → R2 : α ∈ A} is an uncountable family of simple closed
curves in the plane, then there is an element fβ of F and sequences fαn and fβn of
elements of F so that the curves fαn converge to fβ metrically from the exterior of
the simple closed curve fβ while the curves fβn converge to fβ metrically from the
interior of the simple closed curve fβ .
3.12. [Baire Category Theorem] Suppose that C is a nonempty compact metric
space and that C is the union of countably many closed subsets C1 , C2 , . . .. Then
at least one of the sets Cn contains a nonempty open subset of C.
3.13. [The Slicing Theorem] Suppose that C is a compact subset of the disk
D = [0, 1] × [0, 1] ⊂ R2 , and suppose that every horizontal slice [0, 1] × {y} in
the disk D contains a nonempty interval [ay , by ] × {y}. Show that C contains a
nonempty open subset of D so that the dimension of D is 2.
3.14. Can the Slicing Theorem be generalized to compact subsets of the 3-
dimensional cube?
CHAPTER 4
We will prove the existence of a partition of unity at the end of this section.
But for purposes of this proof, we assume the existence. That is, we apply the
existence for the space X = int(B2 ) and the given cover U.
If x ∈ int(B2 ), we define f (x) = fi (x) · xi . Since the support of fi lies in ui ,
fi (x) can only be nonzero for the one or two elements of U that contain x. If x is
in only one element ui , then f (x) = xi . If x is in the two elements ui and uj , then
f (x) = fi (x) · xi + fj (x) · xj , with fi (x), fj (x) ≥ 0 and fi (x) + fj (x) = 1. That is,
f (x) is a convex combination of xi and xj , hence lies in the interval [xi , xj ] ∈ X.
Since the cover U is locally finite, f (x) is locally a finite sum of continuous
functions, hence continuous.
Since the diameters of the sets ui , and also of unions ui ∪ uj with ui and uj
intersecting, go to 0 near ∂B2 , the map f extends continuously to the boundary of
B2 by the identity on S1 = ∂B2 .
This completes the proof.
The metric version of Urysohn’s Lemma is adequate for our purpose here. We
recall its statement.
Lemma 4.8 (Urysohn’s Lemma, metric case). Suppose that X is a metric space
and that A and B are disjoint closed subsets of X. Then there is a continuous
function f : X → [0, 1] such that f (A) = {0} and f (B) = {1}.
Proof of the existence of a partition of unity. We are given a locally
finite open cover U = {u1 , u2 , . . .} of a separable metric space X, and we seek a
partition of unity subordinate to the cover U.
By the shrinking lemma, there is an open cover V = {v1 , v2 , . . .} of X such
that, for each i, the closure of the set vi lies in ui . By Urysohn’s Lemma, there is
a continuous function gi : X → [0, 1] that is 0 on the complement of ui and is 1 on
4.4. TECHNIQUES NEEDED IN HIGHER DIMENSIONS 41
4.5. Exercises
4.1. Solve Exercise 4.3 on page 37.
Solution. For each positive integer n, let Cn = {x ∈ Rn | |x| ∈ [n − 1, n]}.
The sets Cn are compact. Hence there is a finite subset Un of U that covers Cn .
from Un any elements that do not actually intersect Cn . Then we
We may remove
may set V = n Un .
4.2. Solve Exercise 4.9 on page 41.
4.3. Solve Exercise 4.10 on page 41.
4.5. EXERCISES 43
intuitive understanding of the constructions does not require those details. There-
fore, we will describe the constructions without complete justification, then include
the topological and analytical lemmas and their proofs at the end of the section.
Example 5.1 (Pólya’s example). See Figure 1. Begin with a triangle T = abc
having two distinguished vertices a and b. Choose a point c ∈ int(ab). Let C
denote a triangle in T that has c as one vertex and whose opposite side is a tiny
neighborhood N of c in int(ab). Remove int(N ) ∪ int(C) from T so as to form two
new, smaller triangles T0 and T1 , the first having a and c as distinguished vertices,
the second having c and b as distinguished vertices. (See Figure 1.) In the same
manner, cut these two triangles into four triangles T00 , T01 , T10 , and T11 . Iterate,
making sure that, in the limit, triangle size approaches 0. Stage 0 consists of the
single triangle A0 = T . Stage 1 is the union A1 of the two triangles T0 and T1 . Stage
n
2 is the union A2 of four triangles. In general,
∞ An consists of 2 triangles, strung
together linearly. The intersection A = n=0 An , as an intersection of compact,
connected subspaces of T , is also compact and connected. Since the triangles of
stage n are strung together in a linear fashion and since triangle size approaches 0,
the intersection A is obviously linearly ordered with the linear-order topology. It
follows that A is an arc.
T
T0
T1
T00
T01
T11
T10
Q1 Q2
Q0 Q3
The three arcs that we have described are interesting for many reasons:
First, all three arcs are incredibly twisted and kinked. In particular, they
are fractal arcs in the intuitive sense of Benoit Mandelbrot. See Mandelbrot [86],
Devaney [88], and Falconer [87].
Second, we see even more precisely that these arcs can be constructed to satisfy
the technical definition of fractal as described by Mandelbrot, in that they have 2-
dimensional area > 0 while their topological dimension is 1. We see this by noting
that the amount of area removed at each stage can be made as small as desired.
Hence, we may assume that the total area removed in the process can be made as
small as we like: < for any prescribed > 0. Thus the 1-dimensional arc can be
chosen to have positive 2-dimensional area greater than (original area - ). That
is, 1-dimensional sets can have positive 2-dimensional area.
Third, we can string such arcs of positive area together to form a simple closed
curve S. The curve S will separate the plane into two pieces, an interior U and an
exterior V . The closure of U will be U ∪ S. The area of U will be strictly less than
the area of its closure. That is, the boundary of a set can be fat.
Fourth, the edges of the gaps that we cut in creating the arcs can be pinched
back together to recover the original triangle and quadrilaterals (See Figure 4). The
result maps our kinky arcs continuously onto the original triangle and quadrilater-
als. That is, we obtain space-filling curves — first a triangle-filling curve, then two
quadrilateral-filling curves. Figure 5 shows approximations to the quadrilateral-
filling curves.
48 5. FAT CURVES AND PEANO CURVES
Fifth, each of the three space-filling curves we have just described is famous
in its own right. The third curve is the original curve described by Peano, for
which Peano curves were named; Peano described the curve by formula depending
essentially on a ternary representation of the real numbers. The second is Hilbert’s
space-filling curve for which Hilbert drew several approximations and gave the first
visual description of a space-filling curve; Hilbert’s curve is essentially based on
expressing numbers in base 4. The first of the three examples is the most recent. It
is Pólya’s space-filling curve which depends rather obviously on binary expansions
of the real numbers as indicated by the subscripts on the labels of the triangles Tij .
Pólya described the image of each point x ∈ [0, 1] under his space-filling curve
algorithmically. Lax [85] made that algorithm particularly visual as follows: We
assume that x is expressed in binary notation as an unending string of 0’s and 1’s.
x = .10 · · ·
We take as the triangle a right triangle whose legs are not equal in length.
We draw the altitude P0 to the hypotenuse. This path divides the triangle into
two similar right triangles, one of which is smaller and one of which is larger. If the
first entry in the binary expansion of x is 1 (as it is in the indicated example), we
draw the altitude P1 in the larger triangle. If the first entry in the binary expansion
of x is 0, we draw the altitude P1 in the smaller triangle. In either case, the path
5.2. THE TOPOLOGICAL LEMMAS 49
P0
P1
P2
x = .10 · · ·
P1 divides a triangle into two similar triangles, one of which is smaller and one of
which is larger. If the second entry in x is 0 (as it is in the indicated example of
Figure 6), we draw the altitude P2 in the smaller triangle. If the second entry in
x is 1, we draw P2 in the larger triangle. We iterate, considering successive entries
in the binary expansion of x. The sequence P0 , P1 , P2 , . . . converges to a unique
point of the triangle. This convergence point is the image of x under Pólya’s map,
as described by Lax. Note that the paths described by .0111111 · · · and .10000 · · · ,
both of which describe the real number 1/2, converge to the same point of the
triangle, namely the vertex between the two original legs of the triangle. Although
some numbers have two binary expansions, the algorithm gives the same image
point for each.
The examples show that there are unexpected difficulties in defining dimen-
sion. One cannot say that a set is 1-dimensional if it can be parametrized by a
single variable. In fact, the unit cube of dimension n can be parametrized by a
single variable in every dimension n. One cannot say that a set is n-dimensional
topologically just because it has positive n-dimensional volume.
The examples show that the size of a set and the size of its closure may be
different. This causes technical difficulties in assigning size (volume) to a set.
All of these difficulties caused headaches for an entire generation of mathemati-
cians.
that f (yn ) = xn . Since [0, 1] is compact, we may assume that yn → y ∈ [0, 1]. By
continuity, f (y) = x. Therefore, y lies in the open set f −1 (U ). Since [0, 1] is locally
connected, the component V of f −1 (U ) that contains y is open in [0, 1], hence
contains all but finitely many of the points yn . The image of V is a connected
subset of U and contains x. Hence f (V ) ⊂ V , and all but finitely many of the
points xn = f (yn ) lie in V , a contradiction. We conclude that V is open and that
X is locally connected.
Remark. Our proof that X is locally connected used only the facts that [0, 1]
is compact and locally connected. End remark.
Conversely, assume that X is locally connected. We must construct a continu-
ous surjection f : [0, 1] → X. For each integer n > 0, let U(n) be a finite open cover
of X by connected open subsets of diameter < 1/2n . Let V(n) be the collection of
closures of the elements of U(n). Note that the elements of V(n) are also connected
of diameter < 1/2n .
There is a sequence V(1, 1), V(1, 2), V(1, 3), . . . , V(1, n(1)) of elements of V(1)
such that:
(1) Every element of V(1) appears at least once in the sequence, and
(2) For each i > 1, V(1, i) ∩ V(1, i − 1) is nonempty.
Pick points x(1, 1) ∈ V(1, 1), x(1, 2) ∈ V(1, 2) ∩ V(1, 1), . . . , x(1, n(1)) ∈
V(1, n(1)) ∩ V(1, n(1) − 1), and x(1, n(1) + 1) ∈ V(1, n(1)). Let
0 = y(1, 1) < y(1, 2) < · · · < y(1, n(1)) < y(1, n(1) + 1) = 1
denote equally spaced points in the interval [0, 1]. We will eventually map y(1, i)
to x(1, i) for each i.
We proceed by induction. We assume that a sequence V(k, 1), V(k, 2), . . . ,
V(k, n(k)) of elements of V(k) have been defined such that:
(1) Every element of V(k) appears at least once in the sequence.
(2) For each i > 1, V(k, i) ∩ V(k, i − 1) is nonempty.
We assume further that points x(k, 1) ∈ V(k, 1), x(k, 2) ∈ V(k, 2) ∩ V(k, 1),
V(k, 3) ∈ V(k, 3)∩V(k, 2), . . ., x(k, n(k)) ∈ V(k, n(k))∩V(k, n(k)−1)), and x(k, n(k)+
1) have been chosen, along with corresponding points y(k, i) ∈ [0, 1] which are to
be mapped to the points x(k, i).
54 5. FAT CURVES AND PEANO CURVES
We concentrate on a single one of the sets V(k, i) with its special points x(k, i)
and x(k, i + 1) and corresponding points y(k, i) and y(k, i + 1) of [0, 1]. There is
a finite subcollection of the cover U(k) that irreducibly covers V(k, i). (That is,
no smaller subcollection covers V(k, i).) The elements of this subcollection can be
placed in a sequence such that:
(1) Each element appears at least once in the sequence.
(2) Each element intersect the next element in the sequence.
(3) x(k, i) lies in the first element in the sequence.
(4) x(k, i + 1) lies in the last element in the sequence.
Choose corresponding x’s in successive intersections and equally spaced y’s in
the interval [y(k, i), y(k, i + 1)] ⊂ [0, 1]. Use the element x(k, i) as the first of the
x’s and x(k, i + 1) as the last of the x’s. Arrange these covers of the V(k, i)’s into
a single sequence with the corresponding sequence of x’s and y’s.
Iterate. Then the map sending y(k, i)’s to the x(k, i)’s is defined on a dense sub-
set of [0, 1] and is uniformly continuous. It follows that there is a unique continuous
extension to all of [0, 1], and this extension is surjective.
Corollary 5.23. If X is a Peano continuum, U is a connected open subset
of X, and x, y ∈ U , then there is a continuous function f : [0, 1] → U such that
f (0) = x and f (1) = y.
Proof. Cover U with connected open sets whose closures lie in U . There is
a sequence of these open sets such that x lies in the first and y in the last, while
each open set intersects the next one in the sequence. Let X1 , X2 , . . ., Xn denote
the corresponding sequence of closures. Let this chain replace the first chain in the
proof of the previous theorem. Proceed thereafter as in the proof of the previous
theorem. The result is a continuous function f : [0, 1] → U with f (0) = x and
f (1) = y.
Theorem 5.24. Suppose that f : [0, 1] → X is a continuous function into a
metric space X, with f (0) = f (1). Then there is an arc g : [0, 1] → f ([0, 1]) that
joins f (0) with f (1).
Proof. Let I(1) = [a(1), b(1)] denote a largest metric interval in [0, 1] such
that f (a(1)) = f (b(1)). Let f1 (x) = f (x) for each x ∈ / I(1), and let f1 (I(1)) =
f (a(1)). Let I(2) = [a(2), b(2)] denote a largest metric interval in [0, 1] \ I(1) such
that f (a(2)) = f (b(2)). Change f1 only on I2 so as to define f2 with f2 (I2 ) =
f1 (a(2)). Iterate. Since there can only be finitely many subintervals of any positive
size, this process creates a limit function g such that fn → g, g : [0, 1] → f ([0, 1]),
and every point preimage is closed and connected. Thus g([0, 1]) is an arc by
Corollary 5.7.
5.5. Exercises
5.1. Construct a physical model of Pólya’s triangle-filling curve out of two
colors of poster board, each cut into triangles, attached together in a line. Show
how that line can be folded into one large triangle, so that the folding approximates
Pólya’s map. See Figures 8 and 9.
5.2. Construct physical models for the Peano and Hilbert space-filling curves.
5.3. Pólya’s curve can be modified by basing it on right triangles of various
shapes. What shape or shapes minimize the number of points sent to the points of
the large triangle? (See Lax [85].
5.5. EXERCISES 55
The characterization of the simple closed curve S will be reduced to the second
characterization of the arc by showing that any two points a and b of S divide S
into two subspaces, each being an arc with a and b as endpoints.
Before we begin the proofs of the characterization theorems, we need two facts.
The first is very easy. The second is the most complicated ingredient in the char-
acterization theorems.
Lemma 6.4. Suppose that X is a compact, connected metric space, that x ∈ X,
and that X \ {x} = A ∪ B, separated. Then each of the subspaces A ∪ {x} and
B ∪ {x} is connected.
Proof. Suppose to the contrary that A ∪ {x} = C ∪ D, separated, with x ∈ C.
Then X = (B ∪ C) ∪ D, separated, so that X is not connected, a contradiction.
B0
Bn
B∞
B
X x0 xn x∞ y
A0
An
A∞
A
there is a linear order on X with distinct first and last points with re-
spect to which X has the linear-order topology. From this point on, we
assume the hypotheses of the second characterization of the arc: X is a
compact, connected metric space with exactly two points a and b that
do not separate X.
Lemma 6.7. If X \ {z0 } = A ∪ B, separated, with a ∈ A, then b ∈ B.
Proof. By Lemma 1.6, each of A and B must contain a point that does not
separate X. Since A contains a, B must contain the other nonseparating point
b.
Definition 6.8. For each point z ∈ X, we define sets A(z) and B(z) as follows.
If z ∈ A \ {a, b}, then, by hypothesis, A \ {z} = A(z) ∪ B(z), separated, with
a ∈ A(z). By the previous lemma, b ∈ B(z). For the nonseparating point a, we
define A(a) = ∅ and B(a) = A \ {a}. Similarly, A(b) = A \ {b} and B(b) = ∅. We
see that these sets are uniquely defined as follows.
Lemma 6.9. The sets A(z) and B(z) are connected, hence uniquely defined.
Proof. If B(z) = C ∪D, separated, with b ∈ C, then X \{z} = (A(z)∪C)∪D,
separated, so that D contains a third nonseparating point of X, a contradiction.
Definition 6.10. We define x < y in X if x ∈ A(y).
Lemma 6.11. The relation x < y is satisfied iff y ∈ B(x). It is impossible to
have both x < y and y < x. The relation < is also transitive. Therefore < is a
linear order on X.
Proof. We have x < y iff x ∈ A(y). In that case, the connected set {y} ∪ B(y)
misses x and contains both y and b. Hence {y} ∪ B(y) ⊂ B(x) and, in particular,
y ∈ B(x). Conversely, if y ∈ B(x), then the connected set {x} ∪ A(x) misses y and
contains both x and a. Hence{x} ∪ A(x) ⊂ A(y) and, in particular, x ∈ A(y) and
x < y.
To have x < y and y < x we would have to have y ∈ B(x) by the previous
paragraph, and y ∈ A(x) by definition. But B(x) and A(x) are disjoint.
The order is transitive because x < y < z implies that {x}∪A(x) ⊂ {y}∪A(y) ⊂
A(z), so that the order is transitive, hence a linear order.
Lemma 6.12. If X is a compact, connected metric space with exactly two non-
separating points a and b, then the linear-order topology on X defined by the linear
order above agrees with the given metric topology on X.
Proof. Intervals of the form [a, y) = A(y) and (x, b] = B(x) are open and
form a subbasis for the linear-order topology on X. Hence the map from the metric
topology to the linear-order topology on X is continuous.
We need to show that, conversely, if z is a point of the open set U in the metric
topology, then z is contained in an open interval in U in the linear-order topology.
Suppose z ∈ U is distinct from the initial endpoint a. Since the set {z} ∪ A(z)
is connected, there is a sequence z1 < z2 < · · · from A(z) converging to z in the
metric topology. We claim that, for some i, the order interval (zi , z) lies in U .
If not, then we may assume the existence of points w1 , w2 , . . . ∈ A(z) \ U with
z1 < w1 < z2 < w2 < · · · such that w1 , w2 , · · · → w ∈ A(z) \ U . Choose w with
w < w < z. Since zi → z, we may assume that all of the zi ’s lie in the open
6.2. THE CANTOR SET AND ITS CHARACTERIZATION 61
set B(w ). It follows that the wi ’s lie in B(w ) as well. But the wi ’s converge to
w ∈ A(w ), a contradiction.
By a similar argument, there is an interval (z, zi ) ⊂ U .
This last lemma completed the final requirement for the application
of first characterization. Hence the proof of second characterization of
the arc is complete.
We now prove three lemmas that, together, establish the character-
ization of the simple closed curve. From here on we assume that S is
a compact, connected metric space having at least two points that is
separated by each pair of its points.
Lemma 6.13. No point of S separates S.
Proof. If S \ {x} = A ∪ B, separated, then A ∪ {x} and B ∪ {x} are connected,
and each has a nonseparating point distinct fom x, say a ∈ A and b ∈ B. Then
S \ {a, b} = (A ∪ {x} \ {a}) ∪ (B ∪ {x} \ {b}) is the union of two connected sets with
a common point, hence is connected, a contradiction.
Lemma 6.14. If S \ {x, y} = A ∪ B, separated, then each of the sets A ∪ {x, y}
and B ∪ {x, y} is connected.
Proof. If A ∪ {x, y} = C ∪ D, separated, we consider two cases.
Case 1. Suppose x and y are in the same one of the sets, say x, y ∈ C. Then
S = (B ∪ C) ∪ D, separated, a contradiction.
Case 2. Suppose x ∈ C and y ∈ D. Then S \ {x} = (C \ {x}) ∪ (B ∪ D),
separated, a contradiction.
Lemma 6.15. If S \ {x, y} = A ∪ B, separated, then each of the sets A ∪ {x, y}
and B ∪ {x, y} is an arc with x and y as endpoints.
Proof. Suppose to the contrary that b ∈ B does not separate B ∪ {x, y}. If
a ∈ A does not separate A ∪ {x, y}, then S \ {a, b} is the union of two connected
sets with x, y as common points, a contradiction. Hence every point of A separates
A∪{x, y} so that A∪{x, y} is an arc with endpoints x and y by the characterization
of the arc. But thus S \{a, b} is the union of the connected set C = (B ∪{x, y}\{b})
and the two components of A ∪ {x, y} \ {a}, each of which shares a point with C,
a contradiction. Hence every point of B separates B ∪ {x, y}, so that B ∪ {x, y} is
also an arc with endpoints x and y.
The characterization of the simple closed curve is an immediate con-
sequence of this last lemma. Hence the characterization is complete.
C0
C1
C2
C3
C4
C (n) (in ) from C whose intersection is a single point c(x) ∈ C. The map x
→ c(x)
is the desired homeomorphism from X to C.
Proof of the first lemma. Consider an arbitrary point x ∈ X. We claim
that x has arbitrarily small neighborhoods in X that are both open and closed in
X. If this is not the case we shall show the component containing x has more than
one point.
In order to see this, we consider two positive numbers and δ. Let N (δ) denote
the set of points y that can be joined to x by a chain y = y0 , . . . , yn = x of points
such that for i = 1, . . . , n, the distance from yi to yi−1 is less than δ. We call such
a chain a δ chain from y to x. Note that the set N (δ) is both open and closed.
We claim that, for δ sufficiently small, the set N (δ) lies in the neighborhdood
6.3. INTERESTING CANTOR SETS 63
k1
We set X1 = j=1 X1 (j) and proceed by induction. Assuming that Xi has
been defined, find a collection of rectangles in each rectangle of Xi that block the
bottom from the top, each of the new rectangles having diameter less than 1/(i+1).
∞ rectangles is Xi+1 .
The union of these smaller
The intersection i=0 Xi is an opaque Cantor set in the sense that every seg-
ment from the bottom of X0 to the top must intersect the Cantor set.
6.3.2. A Fat Cantor Set. The total length of the segments removed from
[0, 1] in forming the middle-thirds Cantor set is
1 1 2 1 2 2
+ · + · · + · · · = 1.
3 3 3 3 3 3
We conclude that the middle-thirds Cantor set has length 1 − 1 = 0.
But the only critical condition required in the construction of the general Cantor
set is that the pieces remaining at each stage have diameters approaching 0 in the
limit. In order to satisfy that requirement, it suffices, for example, to make sure
that each remaining piece have diameter less than one-half the diameter of the
preceding pieces. Thus we may remove a total length as little as we please. As a
result, there are Cantor sets in [0, 1] with length as close to 1 as we choose.
n
2
s
J (C, Un ) = {|u| : u ∈ Un } =
s
= en(log 2−s log 3) ,
3s
where we have used the identity ab = eb log a .
If log 2 − s log 3 = 0, which implies s = log 2/ log 3, we have H s (C, Un ) =
en·0
= 1, which is neither ∞ nor 0. For s < log 2/ log 3, H s (C, Un ) → ∞. For
x > log 2/ log 3, H s (C, Un ) → 0. We conclude that s = (log 2/ log 3 = r) is the
most probable dimension of C In view of the previous theorem, it remains only to
show that H r (C) = 1.
The estimates that we have already made show that H r (C) ≤ 1. It remains to
prove that H r (C) ≥ 1.
To that end, we suppose to the contrary that H r (C) < 1. We fix a positive
number δ, and consider an open cover U of C by sets of diameter < δ such that
H r (C, U ) < 1. Since C is compact, we may assume that the cover U is finite.
Without increasing the sum, we may assume that each element of U is an open
interval. Expanding each element of U a tiny bit if necessary, we may assume that
the end points of the interval lie in the complement of C. We replace these open
intervals by their corresponding closed intervals, and proceed to modify this cover
by closed intervals as follows.
If I = [a, b] is one of the intervals, we may shorten I until its endpoints a and
b actually lie in C. Let E denote the largest gap in C between the endpoints of I.
This gap divides I into three subintervals P , E, and P , with a ∈ P and b ∈ P .
We need three facts:
(1) 3r = 2. (Take the logarithm of both sides.)
(2) (3/2)(|P | + |P |) ≤ |P | + |E| + |P |. (This follows from the fact that
|P |, |P | ≤ |E|. In order to see that |P |, |P | ≤ |E|, consider the last n ≥ 0 such
that the endpoints a and b of the interval I lie in the same interval J of Cn . Then
E is the middle third of J, P is a subinterval of the first third of J, and P is a
subinterval of the last third of J.)
(3) ((|P | + |P |)/2)r ≥ (|P |r + |P |r )/2. (This follows since r < 1 so that the
function xr is concave. See Figure 5)
f (avg) ≥ avg(f )
f (b)
f ((a + b)/2)
A concave curve
f (a)+f (b)
2
f (a)
a (a + b)/2 b
We recall the identification space or decomposition space whose points are the
elements of X. That is, each element of X, even if it consists of more than one
point of [0, 1], is abstractly considered to be one point of X. We assigned a topology
to X in the usual manner as follows. We considered the function p : [0, 1] → X
6.4. CANTOR SETS IN THE PLANE ARE TAME 69
which assigned to each point x ∈ [0, 1] the equivalence class p(x) = [x] ∈ X which
contains x. We declared a subset U ⊂ X to be an open subset of X iff the union
U ∗ = p−1 (U ) of the elements of U formed an open subset of [0, 1]. Note that,
by this definition, the map p : [0, 1] → X is continuous. We proved the following
earlier.
Theorem 6.23. The decomposition space X is an arc.
Definition 6.24. Let I denote the set of closed intervals [xα , yα ] in [0, 1] that
are closures of the open intervals deleted from [0, 1] in defining the middle-third
Cantor set C. Let X denote the partition of [0, 1] into the elements of I and the
singleton sets containing the points not included in any of the intervals of I. We
assign X the identification-space topology. Let p : [0, 1] → X denote the continuous
function that assigns each point x ∈ [0, 1] the element p(x) = [x] ∈ X which contains
x. Note that X is an arc by the previous theorem. Hence we may identify X with
[0, 1] and p with a map from [0, 1] to itself which fixes 0 and 1.
Theorem 6.25. Let I, X = [0, 1], and p : [0, 1] → X = [0, 1] be as in the
previous definition. Then we may assume that p(I) is the set of rational numbers
in the open interval (0, 1).
Proof. The set p(I) is a countable dense subset of (0, 1) ⊂ [0, 1] = X. Every
countable dense subset of (0, 1) is embedded in [0, 1] in exactly the same way.
Corollary 6.26. The set of points of the middle-thirds Cantor set C that are
not endpoints of any of the open intervals removed in forming C is homeomorphic
to the set of irrational points between 0 and 1.
Theorem 6.27. There is a (weakly) monotone continuous function f : [0, 1] →
[0, 1], with f (0) = 0 and f (1) = 1 that has derivative 0 almost everywhere.
Proof. Map the first middle third to 1/2. Map the next larger intervals to
1/4 and 3/4. Map the four next intervals to 1/8, 3/8, 5/8, and 7/8, etc. Extend
continuously to the remaining points of [0, 1]. Since the sum of the lengths of the
intervals defining the Cantor set is
1 2 n 1 1
· 1 + 2/3 + 2/3 + · · · + 2/3 + · · · = · = 1,
3 3 1 − (2/3)
the graph is horizontal on a set of measure 1 yet rises from 0 to 1.
Theorem 6.28 (All Cantor sets in the plane are tame). Suppose C is the
middle-thirds Cantor set and f : C → R2 is an embedding of C in the plane. Then
the map f extends to a homeomorphism F : R2 → R2 .
In order to keep the proof of this theorem in reasonable bounds, we will assume
that finite families of disjoint polygonal disks in the plane can be moved around
at will, in the sense that any two such families consisting of the same number of
disks are embedded in the plane in the same way. If each of the disks is a triangle,
this is a fairly easy fact. With some tedium, the general fact can be reduced to the
following two exercises.
Exercise 6.29. If D is a polygonal disk in the plane, then, without introducing
any new vertices, D can be cut into triangles.
Exercise 6.30. If D is a polygonal disk in the plane and N is any open
neighborhood of D, then there is a homeomorphism of the plane fixed outside of N
that takes D to a triangular disk. [Hint: Use the previous exercise to assume that
D is a union of triangles. Inductively, move D into itself so as to make D have one
triangle fewer until there is only a single triangle remaining.]
The first step in proving this theorem is to prove that f (C) can be covered by
a collection of disjoint polygonal disks in the plane, each of small diameter. This
fact follows from the following more general theorem.
Theorem 6.31. Suppose that K is a compact subset of the plane and that > 0.
Then there is a finite collection D1 , . . . , Dk of disks-with-holes in the plane whose
interiors cover K, each having polygonal boundary, such that, for each i, Di lies in
the neighborhood of some component of K.
Proof. Fix a positive number δ. For each x ∈ K, let E(x) denote a polygonal
disk of diameter < δ whose interior contains x. Since K is compact, there is a finite
collection E(x1 ), . . . , E(x ) whose interiors already cover K. Each can be expanded
slightly so that their boundaries are in general position. (General position means
6.4. CANTOR SETS IN THE PLANE ARE TAME 71
in this case that two boundaries intersect, if they intersect at all, in finitely many
points, each in the interior of edges that cross one another at those points.) Then
the union Kδ = i=1 E(xi ) is a finite collection of disks-with-holes whose interiors
cover K, each having polygonal boundary.
We will show that, for some positive integer n, K1/n satisfies the conclusions of
the theorem. Suppose not. Then for each n > 0, there is a component Cn of K1/n
that is not in the neighborhood of any component of K. Passing to a subsequence
of C1 , C2 , . . . if necessary, we may assume that the sequence C1 , C2 , . . . converges
to a compact connected set C that necessarily is a subset of K since every point
of Cn is within 1/n of a point of K. See Theorem 3.16. Let N denote the
neighborhood of C. Then, for n sufficiently large, each of the sets Cn must lie in
N ; otherwise, again passing to a subsequence if necessary, we may assume that
there is a convergent sequence x1 , x2 , . . ., with xn ∈ Cn \ N , which converges to a
point x in the lim sup of the Cn ’s but not in C, a contradiction. This completes
the proof of the theorem.
Proof. Choose δ > 0 so small that any disk containing a point of some f (Ci )
cannot contain a point of another f (Cj ). Also choose δ > 0 so small that the δ
neighborhood of f (Ci ) still lies in an xy-square of edge length < . Use the Covering
Lemma to cover f (C) by disjoint disks of diameter < δ. Then these disks fall into
finite subcollections, namely, those containing a point of f (C1 ), those containing a
point of f (C2 ), etc. These subcollections can then be consolidated by the Clumping
Lemma 6.33 into single disks Di for each subcollection.
Proof that all Cantor sets in the plane are tame. The middle-third
Cantor set C is naturally covered by a collection I(n) of 2n intervals of length 1/3n
for every n. See Figure 9. Each of those collections of intervals can be thickened
slightly to create coverings D(n) by disks.
Assuming that D(nj ) has been chosen and that nj+1 > nj , for each interval
I ∈ I(nj+1 ), we cover f (C ∩ I) by the interior of a polygonal disk D(I). We may
choose these disks to be disjoint; and, if nj+1 is sufficiently large, we may assume
that the diameter of each D(I) is < j + 1 and that D(I) lies in the interior of a
disk element of D(nj ). Let the collection of these disjoint disks be D(nj+1 ).
Disjoint polygonal disks A1 , . . . , Ak in the interior of a disk A can be moved, in
order, to any other collection B1 , . . . , Bk in the interior of A by a homeomorphism
of A that fixes the boundary of A. Using this fact, we move the disks of D(n1 ) to
the disks of D(n1 ) by a homeomorphism of R2 . Then, fixing all points outside the
union of these disks, we can move the images of the disks of D(n2 ) to the disks of
D(n2 ). In general, fixing all points outside the union of the disks of D(nj ), it is
possible to move the images of the disks of D(nj+1 ) to the disks of D(nj+1 ). The
limit of these homeomorphisms is a homeomorphism g : R2 → R2 that, restricted
to f (C) is the inverse of f . This completes the proof of the theorem.
6.5. Exercises
6.1. Solve Exercise 6.20 on page 64.
6.2. Solve Exercise 6.29 on page 70.
6.3. Solve Exercise 6.30 on page 70.
6.4. The points of the middle-thirds Cantor set can be expressed explicitly as
infinite expansions using ternary numbers (symbols using only 0, 1, and 2). How
is that done? Likewise the points of the interval [0, 1] can be expressed as infinite
binary expansions. How is that done? Use these expansions to define an explicit
mapping from the middle-thirds Cantor set onto the interval [0, 1]. This, of course,
shows that there are as many points in the middle-thirds Cantor set as in the
interval [0.1].
6.5. How can you define different Cantor sets with different Hausdorff dimen-
sions?
CHAPTER 7
Algebraic Topology
We claim that there is at least one half [a , b ] of the interval [a, b] such that z
represents a nonzero element of Hj (Sn \ B[a , b ]).
To that end, we let c denote the midpoint c = (a + b)/2 of the interval [a, b].
We set U = Sn \ B[a, c] and V = Sn \ B[c, b]. Then U ∩ V = Sn \ B[a, b] and
U ∪ V = Sn \ B[c]. We examine a segment of the reduced Mayer-Vietoris sequence
of the pair (U, V ) (Fact 3, section 7.1):
Hj+1 (U ∪ V ) → Hj (U ∩ V ) → Hj (U ) ⊕ Hj (V ) → Hj (U ∪ V ).
Since U ∪V = Sn \B[c] the inductive hypothesis implies that the first and last terms
of the segment are equal to 0. Hence, the central homomorphism is an isomorphism,
induced by inclusion of cycles. In particular, z represents a nonzero element either
of Hj (U ) or of Hj (V ), as claimed.
We conclude, inductively, that there are intervals [a1 , b1 ] = [0, 1], [a2 , b2 ],
[a3 , b3 ], . . ., each equal to one-half of the preceding, such that z represents a nonzero
element of Hj (Sn \ B[ai , bi ]) for each i = 1, 2, . . . . Let a denote the intersection of
these intervals. But B[a] is an embedding of the (k − 1)-ball in Sn so that, by
induction, z can only represent the trivial element of Hj (Sn \ B[a]). Hence, there is
a (j + 1)-dimensional chain c in Sn \ B[a] with boundary z. Since the chain c misses
B[a], it also misses all but finitely many of the sets B[ai , bi ]. But that contradicts
the fact that z represents a nontrivial element of Hj (Sn \ B[ai , bi ]).
We conclude that no such cycle z can exist, so that H∗ (Sn \B[0, 1]) is trivial.
does not separate S2 , there is an arc A(U ) that connects a point u ∈ U irreducibly
to f (S1 ) but misses A. (That is, only the endpoints of A(U ) lie in {u} ∪ f (S1 ).)
The arc A(U ) lies, except for its terminal endpoint u in U , and u ∈ B. That is,
u ∈ ∂U . Thus U ∩ N is not empty, so that x is in the boundary of U . Similarly,
x is in the boundary of V . We conclude that f (S1 ) is the boundary of both U and
V.
CHAPTER 8
A
U
J ∪ U and J ∪ V are disks. The proof that each of the sets J ∪ U and J ∪ V is a
disk will require a number of important separation lemmas, which will occupy the
majority of this chapter. End remark.
Proof that X contains a simple closed curve. Since X has more than
one point and, as a Peano continuum, is arcwise connected, we see that X contains
an arc A joining two points x and y. Let B denote a subarc of A in the interior of
A. Since no arc separates, B does not separate x from y. Hence, since components
of open sets in a Peano continuum are arcwise connected, there is an arc C joining
x and y that does not intersect B. The arc C contains a subarc D that irreducibly
joins one of the two components of A \ B with the other component. The arc A
contains a subarc E that joins the endpoints of D. Then J = D ∪ E is a simple
closed curve.
Our first lemma shows how to create spanning arcs with controlled endpoints.
Lemma 8.5 (Spanning-Arc Lemma). Suppose that J ⊂ X is a simple closed
curve separating X into two components U and V . Suppose that α and β are
disjoint subarcs of J. Then there is an arc A properly embedded in D = J ∪ U with
one endpoint in α and the other in β. (We say that A spans J through U from α
to β.) See Figure 2.
A
U
b1 B b2
J a2
Proof. Suppose to the contrary that the arcs A and B are disjoint. We may
assume that the endpoints appear in the order a1 , b1 , a2 , b2 on J. By the Spanning-
Arc Lemma 2, there is an arc C that spans J through V with endpoint c1 between
b1 and a2 and with endpoint c2 between b2 and a1 .
We obtain a contradiction by showing that the simple closed curve K = A ∪
a2 c1 ∪ C ∪ c2 a1 does not separate X. In fact, we show that every point x in
the complement of K can be joined by an arc in X \ K to the connected set
L = B ∪ (J \ K). If x is not already in L, then x must either lie in U \ K or
in V \ K. If x ∈ U \ K, then there is an arc α irreducibly joining x to J \ K in
the complement of the arc c2 a1 ∪ A ∪ a2 c1 . The arc α must miss K entirely. If
x ∈ V \ K, then there is an arc β irreducibly joining x to J \ K in the complement
of the arc a2 c1 ∪ C ∪ c2 a1 . The arc β must miss K entirely.
We show the strength of the Arc-crossing Lemma by deducing some important
corollaries.
Corollary 8.7 (Theta Curve Theorem). Consider a space Θ that is the union
of three arcs A, B, and C having disjoint interiors but that share their endpoints x
84 8. CHARACTERIZATION OF THE 2-SPHERE
and y. Then the complement X \ Θ has three components, namely U with boundary
A ∪ B, V with boundary B ∪ C, and W with boundary C ∪ A. See Figure 4.
x B y
W
C
Proof. Since it requires two of the arcs A, B, and C to separate X, given any
point z in the complement of Θ, there are two arcs α(z) and β(z) with initial point
z and terminal points α (z) and β (z), each arc irreducibly connecting z to Θ, such
that α (z) and β (z) lie in the interiors of different arcs among the three A, B, and
C. We suppose for purposes of argument that α (z) ∈ A and β (z) ∈ B.
The union α(z) ∪ β(z) contains an arc spanning J = A ∪ B through one of the
components U of X \ (A ∪ B). Since its terminal endpoints α (z) and β (z) separate
the ends x and y of C in the simple closed curve A ∪ B, and since α(z) and β(z)
do not intersect C, it follows from the Arc-crossing Lemma 8.6 that C lies in the
other component U of X \ (A ∪ B).
Since all points of U can be irreducibly joined to Θ by an arc ending in the
interior of C, and since all the points of U can be irreducibly joined to Θ by arcs
ending in the interiors of both in A and B, we conclude that U is characterized by
that property.
We find similarly that the points joinable to both B and C form an open set V
bounded by B ∪ C and that the points joinable to both C and A form an open set
W bounded by C ∪ A. The three sets U , V , and W exhaust the points of X \ Θ,
and the proof of the corollary is complete.
B
C
U
J
A B C
x y z
Figure 6. The Contrary Neighbor Graph
86 8. CHARACTERIZATION OF THE 2-SPHERE
y x z
Proof. Assume to the contrary that the indicated graph can be embedded
in X. See Figure 7. Consider the simple closed curve J = Ax ∪ xC ∪ Cy ∪ yA.
The arcs AzC and xBy would have to be properly embedded arcs whose endpoints
separate one another on J. By the Arc-crossing Lemma 8.6, one of these two arcs
would have to have interior lying in one domain U of X \ J, and the other would
have to have interior lying in the other domain V of X \ J. But the arc Bz would
then have to cross the curve J, which separates B from z, a contradiction.
Corollary 8.10. It is impossible to embed the complete graph K5 on five
vertices in X. See Figure 8.
a b
d c
B
C
use of the Theta Curve Theorem 8.7 which, in application, says that a spanning
arc in D divides D into two domains bounded by simple closed curves.
Definition 8.13. Let X, J, U , and D = J ∪ U be given, as above. A simple
subdivision of D is a tuple (D, α1 , α2 , . . . , αk ) such that α1 is a spanning arc in D
that divides D into two closed domains D0 and D1 , α2 is a spanning arc either
of D0 or of D1 that divides that domain into two closed domains, etc. Thus,
thinking inductively, (D, α1 , . . . , αk−1 ) is a simple subdivision that divides D into
finitely many closed domains, each bounded by a simple closed curve; and αk is a
spanning arc in one of those domains that divides it into two closed domains. The
closed domains into which α1 , . . . , αk divide D are called the cells of the simple
subdivision. The arcs α1 , . . ., αk are called the arcs of the simple subdivision. The
union of these arcs with J is called the 1-skeleton of the subdivision.
Lemma 8.14 (Subdivision Lemma). Given > 0, there is a simple subdivision
(D, α1 , . . . , αk ) of D each of whose cells has diameter < .
Proof. The Separation-Triples Lemma 8.12 states that there is a finite col-
lection (A1 , B1 , C1 ), . . . , (Ak , Bk , C ) of separation triples in D such that, if x and
y are two points of D at distance at least from each other, then there is one of
the separation triples (Ai , Bi , Ci ) such that x is in one of the two sets Bi and Ci ,
while y is in the other.
There is a positive number δ that is smaller than the distance from each of the
arcs Ai to its companion sets Bi and Ci . We now process each of the arcs Ai in
turn. Each will supply us with a finite number of the curves αj .
First, we keep the arc A1 as α1 .
Assume inductively that arcs A1 , . . ., Ai−1 have been processed, yielding a
simple subdivision (D, α1 , . . . , αk(i−1) ) of D. Consider the components of Ai \ (J ∪
α1 ∪ · · · αk(i−1) ) whose endpoints cannot be joined in J ∪ α1 ∪ · · · αk(i−1) by an arc
of length < δ. Since J ∪ α1 ∪ · · · αk(i−1) is locally connected and the components of
Ai \ (J ∪ α1 ∪ · · · αk(i−1) ) are countable in number with diameters approaching 0,
there are only finitely many components whose endpoints cannot be so joined. We
add the closure of each such component to the collection of αj ’s.
8.2. EXERCISES 89
8.2. Exercises
8.1. Show that a space containing either of the graphs in Figure 10 cannot
satisfy the Jordan Curve Theorem.
8.2. Show that each of the two graphs in Figure 10 can be embedded in a
sphere with handles. Does a torus suffice? (One handle.)
8.3. If D denotes the square and A is an arc with its endpoints on opposite
sides of D, then A separates the top and bottom of D. (See Figure 11.) However,
there is a connected subset K of D that contains both sides of D but does not
separate the top of D from the bottom of D. Construct such a set K.
90 8. CHARACTERIZATION OF THE 2-SPHERE
2-Manifolds
9.2. Exercises
The Four-color Theorem says that every map on the sphere can be colored
with four colors in such a way that no two countries share a common edge. (A map
consists of a number of polygonal disks with disjoint interiors, but possibly sharing
parts of their boundary.) This theorem is extremely difficult. The same result is
not true for other surfaces.
91
92 9. 2-MANIFOLDS
a b
b a
9.1. Construct 4 disks in the plane that have disjoint interiors such that each
shares an edge with the other three.
9.2. Show that it is impossible to have 5 disks in the plane that have disjoint
interiors such that each shares an edge with the other four. (This result is a
necessary property in order that the Four-color Theorem be true, but it is not
sufficient. The subtle question is whether some exotic complicated map might force
two countries to share an edge, not because of some local configuration, but because
of some global configuration.)
9.2. EXERCISES 93
9.3. Construct 5 disks in the torus that have disjoint interiors such that each
shares an edge with the other 4.
9.4. On a sphere with n handles, how many disks with disjoint interiors can
share a common edge with each of the others?
CHAPTER 10
Our goal is to understand arcs in all 2-manifolds. The fundamental case is the
arc in the 2-sphere S2 .
Definition 10.1. An embedding f : [0, 1] → Sn is called tame if there is a
homeomorphism h : Sn → Sn such that h ◦ f ([0, 1]) lies in a great circle. That is, f
is equivalent to a standard embedding. Otherwise, the embedding f is called wild
(and similarly for embeddings in Rn ).
Before considering arcs in the general 2-manifold, we show that every arc in
the 2-sphere S2 is tame and that tame arcs can be pushed around at will.
Example 10.2. The diagram illustrates a wild embedding of [0, 1] in R3 .
B
J
J−1
J2
J−2 J1
J0
Have the diameters of the curves Jn approach 0 as n → ±∞ and have the curves
±∞
Jn approach the ends of A as n → ±∞. Then the union n Jn \ A, together with
the endpoints of A, will contain two arcs joining the endpoints of A. The union of
A with one of these arcs will form the desired simple closed curve containing A.
10.2. DISK ISOTOPIES 97
Scholium 10.5. The simple closed curve J ⊃ A may be chosen arbitrarily close
to A.
Corollary 10.6. Every arc A = f ([0, 1]) in the plane R2 locally separates R2
into two components.
The proof of the lemma is essentially local so that we can conclude the following.
Lemma 10.7. If A is an arc in a 2-manifold, then A is a subset of a simple
closed curve in the 2-manifold.
Proof. Let X and Y denote the arcs into which ∂A divides ∂D. Let D1
and D2 denote the two disks into which A divides D, with ∂D1 = A ∪ X and
∂D2 = A ∪ Y . Let E1 and E2 denote the two disks into which B divides D, with
∂E1 = B ∪ X and ∂E2 = B ∪ Y . Let g : A → B denote a homeomorphism
that is the identity on ∂A = ∂B. By the Boundary Extension Theorem, there are
homeomorphisms G1 : D1 → E1 and G2 : D2 → E2 that are equal to the identity
on ∂D and are equal to g on A. Then G = G1 ∪ G2 : D → D is a homeomorphism
that is the identity on ∂D. By Alexander’s trick, Theorem 10.10, there is an isotopy
from the identity to G. This isotopy takes A to B.
Proof. Suppose that a ∈ J cuts J into the open arc α and that, in some
natural linear order on α, points b, c, d appear in the order b < c < d. See Figure 4.
d
a
We have seen that there is a simple closed curve K1 bounding a disk D1 con-
taining the arc abcd ⊂ J as a spanning arc and excluding the complementary arc
da ⊂ J. See Figure 5.
a d
K1
a
d
b D1
b c
c
curve K2 very close to the arc c d a b , with K2 bounding a disk D2 that contains
the arc c d a b as a spanning arc and excludes the arc b c . The arc c d a b can
be replaced by a spanning arc that is a finite union of straight line segments. The
curve J can then be moved to the resulting simple closed curve J that satisfies
the requirements of the lemma.
10.3. Exercises
10.1. Solve Exercise 10.12 on page 98.
10.2. Solve Exercise 10.13 on page 98.
10.3. Remove a disk from a torus to create a hole in the torus. Show how to
turn this punctured torus inside out through the hole.
10.4. In the punctured torus, insert a rope and tie it around the center of the
torus. When the punctured torus is turned inside out, what happens to the rope.
It ends up on the outside of the torus. Does it fall away from the torus?
10.5. As we have seen, there is only one embedding of an arc or simple closed
curve in the 2-sphere S2 up to equivalence. That is, if f, g : [0, 1] → S2 are
embeddings, then there is a homeomorphism h : S2 → S2 such that, for each
x ∈ [0, 1], h(f (x)) = g(x). Construct a graph Γ such that there are two embed-
dings f, g : Γ → S2 for which no such homeomorphism h exists.
S2
Example 11.3. Consider the middle-thirds Cantor C set in the real line R.
Multiply by the unit interval to form a Cantor set of intervals, C × [0, 1] ⊂ R2 ⊂
R2 ∪ {∞} = S2 in the 2-sphere S2 . Collapse each of the intervals {x} × [0, 1] with
x ∈ C to a point to form a decomposition space S2 / ∼ . Then this decomposition
space is homeomorphic to S2 . See Figure 2.
101
102 11. R. L. MOORE’S DECOMPOSITION THEOREM
sets containing p(g1 ) and p(g2 ), respectively. Hence S/G is Hausdorff and condition
(2) is satisfied.
Assume condition (2), the Hausdorff condition. Suppose that F is a closed
subset of S. Then F is compact since S is compact. Hence p(F ) is compact
by continuity. Since S/G is Hausdorff, compact subsets of S/G are also closed.
Therefore p(F ) is closed and condition (3) is satisfied.
Assume condition (3), the continuity condition. Let g be an element of G and
U an open set containing g. Then the complement S \ U is closed, and the image
p(S \ U ) is therefore a closed set C that misses p(g). Thus V = (S/G) \ C is an open
subset of S/G. Hence p−1 (V ) is open and consists of the union of the elements of
G contained in U . Hence condition (1) is satisfied.
q does not separate x from y. A path from x to y in X \ {q} also misses all but
finitely many of the sets In , a contradiction. We conclude that A does not separate
X.
Suppose that J is a simple closed curve in X. Let p1 , p2 ∈ J cut J into two
arcs A1 and A2 . Then A1 and A2 are compact, connected, and have nonconnected
intersection p1 ∪ p2 . The reduced Mayer-Vietoris homology sequence for the pair
U = S2 \ A1 and V = S2 \ A2 contains the segment
H1 (S2 \ A1 ) ⊕ H1 (S2 \ A2 ) → H1 (S2 \ (p1 ∪ p2 )−→H0 (S2 \ J ).
α
The groups H1 (S2 \ A1 ) and H1 (S2 \ A2 ) are = 0 since A1 and A2 are connected;
hence the homomorphism α is one-to one. The group H1 (U ∪ V ) is not 0 since
A1 ∩ A2 is not connected; hence the group H0 (S2 \ J ) is not 0. That is, J separates
106 11. R. L. MOORE’S DECOMPOSITION THEOREM
AW
d
AU
AV
11.4. Exercises
11.1. Solve Exercise 11.9 on page 105.
11.2. Figure 5 indicates three decompositions of the 2-sphere. In the first, the
equator is to be identified to a single point. In the second, the black disk is to be
identified to a single point. In the third, the components of a product of a Cantor
set with an interval are to identified to points. Identify the spaces that result.
12.1. Tools
Our proof of the Open Mapping Theorem depends on the n-dimensional ver-
sions of four theorems that we have already discussed. We have proved all of these
except for the general version of the No Retraction Theorem, and we outlined the
necessary additions that must be carried out to complete the general version.
Theorem 12.3 (Jordan Separation Theorem). If f : Sn → Sn+1 is continuous
and one-to-one, then the image f (Sn ) separates Sn+1 into two components U and
V and is the boundary of each. See Theorem 7.6.
Theorem 12.4 (No Retraction Theorem). There is no continuous function
f : Bn → (Sn−1 = ∂Bn ) such that f (x) = x for each x ∈ Sn−1 . (See Theorem
2.17.)
Lemma 12.5 (Urysohn’s Lemma, easy metric case). If X is a metric space
and A and B are disjoint closed subsets of X, then there is a continuous function
f : X → [0, 1] such that f (A) = {0} and f (B) = {1}. (See Theorem 3.10.)
Lemma 12.6 (Tietze’s Extension Theorem, easier metric case). If X is a metric
space, A is a closed subspace, and f : A → [0, 1] is a continuous function, then there
is a continuous function F : X → [0, 1] such that, for each x ∈ A, F (x) = f (x).
See Theorem 3.14.
Because we have completed the proof of the No Retraction Theorem only in
dimension 2, our proof will be complete only in dimension 2. (However, we recall,
that the reader with some facility working with higher-dimensional simplicial com-
plexes should have no difficulty in generalizing the 2-dimensional proof of the No
Retraction Theorem to higher dimensions. One uses a triangulation of the ball by
very small n-simplexes, uses the forbidden retraction to map this triangulation near
to the boundary of the ball, intersects the ray with the image simplexes, and finds
a one-ended arc in the preimage.)
109
110 12. THE OPEN MAPPING THEOREM
y
x
Lemma 12.8 (Borsuk’s Lemma). Suppose that x and y are points of Sn and
that A is a compact subset of Sn \ {x, y}. If A can be contracted to a point in
Sn \ {x, y}, then x and y lie in the same component of Sn \ A.
(C ∪ A) × 1
U A × [0, 1] U
A × [0, 1]
A × [0, 1]
(C ∪ A) × 0
The image of A ∪ C is indicated in the diagram by a bold line. Then we may set
F (α) = J ◦ K. This shows the existence of F and completes the proof of the
Borsuk Lemma.
112 12. THE OPEN MAPPING THEOREM
12.4. Exercise
12.1. The Open Mapping Theorem is often proved in the differentiable case.
Consult any standard graduate text in analysis for the very interesting proof. It
involves the contraction mapping principle which we have already considered and
proved. See, for example, Serge Lang [41].
CHAPTER 13
Triangulation of 2-Manifolds
(1) α fα (Δ) = X
(2) If two images fα (Δ) and fβ (Δ) intersect, then they do so in the image of a
vertex or the image of an edge of each.
13.2. Tools
We will assume one difficult theorem in the proof, namely, the Schoenflies
Theorem 8.2 on page 81, which we proved in an earlier section. We also assume
the key lemma used in the proof of the Schoenflies Theorem, which we called the
Arc-Crossing Lemma 8.6 on page 82.
Theorem 13.4 (The Schoenflies Theorem). If J is a simple closed curve in
R2 , then J separates R2 into two disjoint open sets, namely, the interior int(J) of
J which has compact closure in R2 and the exterior ext(J) which has noncompact
closure in R2 . The closure of int(J) is homeomorphic to the triangular disk Δ.
(We actually only proved the corresponding result for J ⊂ S2 , of which this
theorem is an easy corollary: the one-point compactification of R2 is S2 so that the
earlier theorem applies.)
Lemma 13.5 (Arc-crossing Lemma). Suppose that J is a simple closed curve
in R2 and that A and B are arcs spanning J across int(J). If the endpoints a1 , a2
of A separate the endpoints b1 , b2 of B on J, then the arcs A and B must intersect.
See Figure 1.
113
114 13. TRIANGULATION OF 2-MANIFOLDS
K B
K
A
v(D)
13.4. Exercises
13.1. Solve Exercise 13.8 on page 115.
[Hint: By our results on disk isotopy, every arc in a 2-manifold looks like every
other. If U is a component of D \ K, trace the boundary of U . It either closes up
as a simple closed curve, or it comes back on itself to form a cut point.]
13.2. Suppose that S is a compact, connected 2-manifold with nonempty
boundary. Show that S can be retracted onto a finite graph.
CHAPTER 14
The Structure Theorem for compact, connected, metric 2-manifolds says that
each is composed from copies of four standard pieces by an operation called con-
nected sum.
The Classification Theorem says that, up to homeomorphism, such 2-manifolds
are completely determined by two invariants, namely orientability and Euler char-
acteristic.
D1 D2
14.2. Edge-pairings
Most students of topology are familiar with the description of a 2-manifold M
as a quotient of a disk D under an edge pairing. See, for example, Figure 2.
b
a a
a
b
v b
Figure 2. The standard presentation of the torus as an identifi-
cation space
d
c
a
c
d b
Figure 3. The disk associated with the word W = abc−1 c−1 b−1 dad
deleting the labels of α and β from W . The two possible situations are pictured in
Figures 4 and 5.
D1
α D2
α
ββ α β
D D
D1
β D2
α β
α
β
α
D D
E E E E
D1 D2 D1 D2 D1 D2 D1 D2
a a a a a a a a
Case 1 Case 2 Case 3 Case 4
In cases 1 and 2, there is nothing in the word W , before the collapse of α with
β to form D , that requires the pinching of D1 with D2 at the vertex in the center
of the picture of E. Hence the diagrams of cases 1 and 2 split at that central vertex
and become pictures of S2 minus a disk and P2 minus a disk, shown in Figures 7
and 8. We conclude that S = D/W is the connected sum of S = D /W with S2
in case 1 and of S = D /W with P2 in case 2:
14.3. PROOF OF THE STRUCTURE THEOREM 121
S 2 \ int(D)
a a
P2 \ int(D)
a a
T 2 \ int(D)
a a
K 2 \ int(D)
a a
That is, the reintroduction of the edges α and β leads to a connected sum of
D /W with one of the four model 2-manifolds. This completes the inductive proof
of the Structure Theorem.
Alternative description. Every surface can be built from the following build-
ing blocks (Figure 11): the 2-dimensional sphere S2 , the annulus A2 , and the Möbius
strip M2 . The 2-sphere is the surface of a ball. The annulus is an open cylinder
with two circles as boundary. The Möbius strip is formed by sewing two ends of a
strip together with a half twist; it has one boundary component.
S2
A2
M2
boundary to boundary. In the figure, the annuli appear as handles on the sphere
and the Möbius bands are indicated by crosses.
Left tfeL
Left Left
Left
n-simplexes are n-cells. The Euler characteristic is the number of even dimensional
cells minus the number of odd dimensional cells.
Some beautiful examples of 2-dimensional cell complexes are the Platonic sur-
faces (the surfaces of the solid tetrahedron, cube, octahedron, dodecahedron, icosa-
hedron) which are 2-spheres tiled by 2-dimensional polygonal disks, with associated
1-dimensional edges and 0-dimensional vertices. See Figure 5 on page 143. In Chap-
ter 16, Euler characteristic will be discussed further.
Remark. By means of algebraic topology, one proves that the Euler charac-
teristic of a cell complex depends only on the topology of the space in question and
not on the particular division into cells. In particular, any triangulation may be
used to calculate the Euler characteristic of a 2-manifold. End remark.
Remark (Motivation for the definition of Euler characteristic). The simplest
equivalence relation on cell-complexes is simple subdivision: cutting one n-cell into
two n-cells by inserting an (n − 1)-cell between them. See Figure 14.
dim n − 1
dim n
dim n
simple subdivision dim n
14.5. Exercises
14.1. Solve Exercise 14.4 on page 119.
14.2. Show that each of the five Platonic surfaces can be obtained by repeated
simple subdivisions from the following cell structure on the 2-sphere which has two
0-cells on the equator, two 1-cells on the equator joining the two 0-cells, and two
2-cells, namely, a northern hemisphere and a southern hemisphere, with resulting
Euler characteristic of 4 − 1 = 2. In particular, all of the Platonic solids have Euler
characteristic equal to 2. This is of course a consequence of the fact that each
is a cellular subdivision of the 2-dimensional sphere and the fact that the Euler
characteristic of a space does not depend on any particular cellulation.
For example, see Figures 15 and 16.
7-5=2
8-6=2
4-2=2 6-4=2
5-3=2
14.3. Show that the model manifold K2 is the connected sum of two copies of
P .
2
[Hint: Triangulate the surfaces M and N so that the disks removed in forming
the connected sum are triangles in the triangulations.]
14.7. Show that every surface M is a connected sum of a sphere with k ≥ 0
copies of T2 and = 0, 1, or 2 copies of P2 . Show that χ(M ) = 2 − 2k − .
14.8. Show that two surfaces are homeomorphic iff they have the same Euler
characteristic and both are orientable or both are nonorientable.
The following nontrivial exercises can be used to show that connected sum is
well-defined in dimension 2.
14.9. Suppose that D1 and D2 are disks in the interior of a disk D. Then show
that there is an isotopy of D, fixing the boundary of D, which takes D1 at stage 0
of the isotopy to D2 at stage 1 of the isotopy.
14.10. Suppose that D1 and D2 are two disks in a connected 2-manifold M .
Then show that there is an isotopy of M that takes D1 at stage 0 to D2 at stage 2
of the isotopy.
14.11. Show that if D is a disk in a 2-manifold M , then there is a homeomor-
phism h : M → M such that h(D) = D and such that h|∂D reverses the orientation
of ∂D.
14.12. Show that if h : S1 → S1 is orientation preserving, then h is isotopic to
the identity map from S1 to itself.
[Hint: Without loss of generality, there is a point x such that h(x) = x. Cut
S1 at x to form an interval (a ball of dimension 1). Apply the Alexander trick.]
14.13. Show that the connected sum M1 #M2 of two compact, connected 2-
manifolds M1 and M2 is well-defined.
A standard way to construct surfaces is to identify the edges of a polygonal
disk in pairs.
14.14. Given the edge pairings from Figures 17 and 18, identify the surfaces as
connected sums of spheres, tori, projective planes and Klein bottles.
14.5. EXERCISES 127
−1
e
c−1
d−1
a−1
a
b−1
c
f −1
a−1
a
j
i
i
e
h−1
c−1
b
h
d f
j
The Torus
The torus T2 is the surface of the doughnut. It is the next simplest surface
after the 2-sphere S2 . Our goal is to understand the simple closed curves on the
torus. A good reference is Rolfsen [96].
As we shall see, the torus T2 is closely related to the Euclidean plane. We begin
our preparations for the study of curves on the torus by examining some delightful
properties of lines and arcs in the plane.
Problem 15.2. Suppose that A = A(0) is an arc joining (0, 0) and (1, 0) in the
plane. Let A(1/2) denote the arc formed by translating A(0) 1/2 unit to the right.
Prove that A(0) and A(1/2) must intersect (Figure 1). See [96].
A = A(0)
A(1)
q
A(1/2)
129
130 15. THE TORUS
Solution. Consider A = A(0) and its two translates A(1/2) and A(1). As-
sume that, contrary to the statement, A(0) does not intersect A(1/2). Then A(1/2)
intersects neither A(0) nor A(1).
Let p ∈ A(1/2) denote a highest point of A(1/2). Let q denote a lowest point
of A(1/2). Form a homeomorphic copy L of the line R by adding to the arc
pq ⊂ A(1/2) a rising vertical ray from p to ∞ and a falling vertical ray from q to
∞. By Exercise 15.1, L separates R2 , and in fact must separate A(0) from A(1).
But A(0) and A(1) intersect at (1, 0), a contradiction.
Exercise 15.3. Let A(0) and A(1) be as above. Let d be any distance between
0 and 1, and let A(d) denote the image of A(0) translated a distance d to the right.
Then show that A(d) must intersect either A(0) or A(1). See Rolfsen [96].
Definition 15.4. If A(0) and A(d) intersect at a point (a, b), then the points
(a, b) and (a − d, b) are points of A(0) at the same height b and at distance d from
each other. We call the subarc of A(0) from (a − d, b) to (a, b) a d-chord in the arc
A(0).
A simple scaling argument then shows that the Exercise 15.3 implies the fol-
lowing corollary.
Corollary 15.5. If A is an arc in the plane that is a d-chord between its
endpoints (that is, has its endpoints at the same height and at distance d from one
another), and if e is any number between 0 and 1, then there is either an e · d-chord
in A or a (1 − e) · d-chord in A. See Rolfsen [96].
Theorem 15.6 (Chord Theorem). Let n denote a positive integer. Then A(0)
must intersect A(1/n). (That is, A(0) must contain a (1/n)-chord.) See Rolfsen
[96] and Figure 2.
A = A(0)
A(1/n)
A(0)
Since translations of the plane preserve distances and lengths, the torus inherits
local distances and lengths, hence a notion of “straight”, from the plane via the
projection p. We call a curve in the torus T2 straight if and only if it lies in the
image of a straight line L under the projection p : R2 → T2 . We shall eventually
show that every simple closed curve J on T2 that does not bound a disk is isotopic
in T2 to a straight simple closed curve J .
Likewise, T2 inherits a notion of angles and slopes from R2 via p since vectors
A and B that are identified have neighborhoods that are identified by a translation
that preserves angles and slopes.
Theorem 15.12 (Straight-curve Theorem). The projection map p : R2 → T2
takes straight line L to a (straight) simple closed curve J ⊂ T2 if and only if the
slope of L is rational. In that case, if we express the slope as a fraction a/b in
lowest terms, take A as a point of L, and set B = A + (b, a), then the segment AB
lies in L, p maps AB onto J, p(A) = p(B), and p is one-to-one on the remainder
of the segment AB.
15.2. THE TORUS AS A EUCLIDEAN SURFACE 133
Proof. Since p preserves distances locally, L has infinite length, and J has
finite length (by compactness), it follows that the map p|L cannot be one-to-one.
If p(A) = p(B) with A, B ∈ L, then A ∼ B, so that there is an integer vector
(b, a) such that B = A + (b, a). But that implies that the slope of L is the rational
number b/a.
Conversely, if the slope of L is rational, then there are points A and B on L
that differ by an integer vector (b, a) so that A and B are identified by p. Fix A ∈ L
and let B denote one of the two points of L \ {A} nearest to A that are identified
with A. Then B − A = (b, a), with b and a relatively prime, so that no two points
of AB other than A and B are identified by p. The slopes of L at A and at B
are, of course, the same; and p preserves those slopes. Hence p(AB) is straight at
p(A) = p(B), so that the image of L entering B extends straight into the image of
AB near A. It follows that p(L) is the simple closed curve p(AB).
134 15. THE TORUS
of Sq. The arcs J and J bound a disk. Expand that disk slightly to form a
disk D such that J and J are spanning arcs of the disk D. We may require that
the extended J is the entire intersection of D with p−1 (J). There is an isotopy
fixed outside of D that moves J to J . This process removes the component J of
p−1 (J) ∩ Sq and joins together the two components that contained the translation
into Sq of the two end pieces of the extended J . This reduces the number of
components of p−1 (J) ∩ Sq, which contradicts our insistence that this number is
already minimal.
15.3. CURVE STRAIGHTENING 135
Sq
D
J
J
We conclude that every component J of Sq ∩ p−1 (J) joins distinct sides of Sq.
Hence, by the Straightening Lemma 10.16, we may assume that each component
J is a straight line interval.
This completes the proof of (2).
Proof of (3). If some component J of p−1 (J) does not cross all horizontal
lattice lines in the same direction, then it crosses some horizontal lattice line twice,
say at points a and b, and intersects no horizontal lattice line between those two
crossings. Because of (2), the interval between a and b in J is monotone in the
horizontal direction. In fact, we claim that the interval from a to b can only consist
of two segments as in the Figure 6.
a b
We assume this claim for the moment. The two segments joining a to b are
drawn as solid segments. Since the set p−1 (J) is invariant under translations by
integer vectors, there are further segments that are translates of the two segments
of ab and are drawn as dotted segments. The dotted segments limit the horizontal
distance between a and b to be less than 1 so that ab acts just like the component
136 15. THE TORUS
J of (2) with respect to a translate of the square Sq. Thus the arc from a to b
can be pushed across the horizontal line to reduce the number of components of
p−1 (J) ∩ Sq, a contradiction. Thus we can complete the proof of (3) by proving
the claim.
If there were a third segment in the path from a to b, we would have the
following situation (Figure 7):
g
f
e
The second edge e has a dotted translate that is forced by its nonintersection
with the first edge to rise higher. The third edge f has a dotted translate that is
forced by its nonintersection with the dotted translate of e to rise even higher, and
so on. The path can therefore never return to the level of a. This proves the claim,
and completes the proof of (3).
Proof of (4). By invariance under translation by integer vectors, p−1 (J)
intersects the top of Sq the same number of times it intersects the bottom, and it
intersects the left side the same number of times it intersects the right side. By
the Edge-adjustment Lemma 10.15, we may assume that the points of p−1 (J) on
any horizontal lattice line are evenly spaced, as are those on any vertical lattice
line. By the Straightening Lemma 10.16, we may assume that all segments are
straight. It follows that the components of p−1 (J) are parallel straight lines, evenly
spaced in R2 . Since the projection is a closed curve, there are points on these lines
that are equivalent under p, hence differ by an integer vector. Hence the slopes are
rational.
Slope = 3/5
15.5. Exercises
15.1. Solve Exercise 15.1 on page 129.
[Hint: Form S2 by adding a point at infinity to R2 . (Remember stereographic
projection.) Apply the Jordan Curve Theorem.]
15.2. Solve Problem 15.2 on page 129.
15.3. Solve Exercise 15.3 on page 130.
15.4. Solve Exercise 15.8 on page 131.
CHAPTER 16
16.1. Orientation
It is surprisingly delicate to define the notions of right and left mathematically.
We considered the notion only by means of the Möbius band in earlier sections.
In general, however, these notions are defined in terms of the determinant. For
2-manifolds, this is done in detail as follows.
Definition 16.1. We defined a 2-manifold as a space X that is locally home-
omorphic with the plane. That is, there are an open cover U = {uα | α ∈ A}
of X and homeomorphisms hα : uα → R2 . These homeomorphisms give rise to
transition functions hβ ◦ h−1
α taking the hα -image of uα ∩ uβ to the hβ -image of the
same set. The open cover U and the accompanying homeomorphisms hα are said
to form an atlas (= a collection of charts or road maps—navigational charts) for
the 2-manifold X.
Definition 16.2. We consider the points of the plane R2 = R×R to be column
vectors (a, b)T . (The superscript T indicates transpose.) We may consider each to
be the tip of an arrow with base at the origin (0, 0)T . We consider a pair (a, b)T and
(c, d)T of such arrows to form a right-handed pair if the determinant with columns
(a, b)T and (c, d)T is positive. That is,
a c
b d = ad − bc > 0.
This is true if, when the right-hand is placed over the origin, then the fingers curl
from the point (a, b)T toward the point (c, d)T .
For example, with standard unit vectors e1 = (1, 0)T and e2 = (0, 1)T (T
denoting transpose),
1 0
|e1 e2 | = = 1 > 0,
0 1
so that the pair is right-handed.
In order to localize this definition, we need to allow vectors to be based at
points (e, f )T not equal to the origin of R2 .
Definition 16.3. We consider points (a, b)T and (c, d)T to form a right-handed
pair with respect to (e, f )T if the determinant with columns (a, b)T − (e, f )T and
(c, d)T − (e, f )T is positive.
139
140 16. ORIENTATION AND EULER CHARACTERISTIC
Take a base point B on the curve J. Pick a center point C on J in the direction
from B of the orientation of the curve J. Then a nearby point R is on the right-hand
side of J if the pair (R, C) is right-handed with respect to B.
Definition 16.7. We say that a path K has intersection number +1 with J if
it crosses J from left-hand side to right-hand side. If K crosses from right to left,
we assign intersection number −1. See Figure 3.
K
+1
cells of dimension n and one new cell of dimension n − 1. These three new cells are
said to subdivide the original cell. See Figure 4.
dim n − 1
dim n
dim n
dim n
Exercise 16.15 (The platonic solids). There are five regular convex polyhedra
whose boundaries consist of faces that are all of the same polygonal shape. See
Figure 5.
Discussion: The polygons meeting at a vertex must have total angle sum at
that vertex of less than 2π. For triangles, this angle restriction dictates that 3, 4, or
5 triangles meet at every vertex. For squares, the only possibility is 3 squares at a
vertex. For pentagons, the only possibility is also 3 at a vertex. These requirements
immediately dictate the combinatorial structure of the polyhedra and lead to the
tetrahedron, the octahedron, and the icosahedron for triangles; the cube for squares;
and the dodecahedron for pentagons.
In a platonic solid, we count the vertices as open 0-cells, the interiors of edges
as open 1-cells, and the open polygons as open 2-cells. Counting vertices, edges,
and faces of each of the Platonic solids yields the following data:
Polyhedron Vertices Edges Faces
Tetrahedron 4 6 4
Octahedron 6 12 8
Icosahedron 12 30 20
Cube 8 12 6
Dodecahedron 20 30 12
In each case,
χ = vertices − edges + faces = 2.
This number χ is called the Euler characteristic of the surface of the polyhedron.
It is, in fact, completely natural to expect that the Euler characteristic of each of
these surfaces should be the same because, combinatorially, each can be derived by
simple subdivision from the same cell complex, as we shall now explain.
Setting. Let Γ denote a finite graph that divides the 2-sphere S2 into disks in
the sense that each component of S2 \ Γ is the interior of a disk whose boundary
144 16. ORIENTATION AND EULER CHARACTERISTIC
is a simple closed curve carried by a subgraph of Γ. The vertices of the graph are
called the open 0-cells of the complex. The interiors of the edges of the graph are
called the open 1-cells of the complex. The components of S2 \ Γ are called the open
2-cells of the complex. There are two sorts of simple subdivision: (1) insertion of
a new vertex into the interior of an edge of the graph. (2) Insertion of a spanning
arc through one of the disks of the complex whose boundary vertices are already
vertices of the graph. The Theta-curve Theorem 8.7 implies that simple subdivision
satisfies the requirements of the setting leading to Euler characteristic.
Exercise 16.16. Let Γ denote a graph in the 2-sphere S2 that has one vertex
and one edge. Let C denote the resulting complex. The Euler characteristic is
therefore χ(C) = 1 − 1 + 2 = 2. If P is a platonic solid, then up to homeomorphism,
P can be derived from C by simple subdivision. Consequently, since χ(C) = 2,
χ(P ) is also 2. If we actually consider the 3-dimensional solid which includes the
interior of the 2-sphere as an open 3-cell, then the Euler characteristic of this solid
is 2 − 1 = 1.
Definition 16.17. The natural setting for the Euler characteristic is the finite
CW-complex C. This is a space formed as the identification space of a finite number
of finite-dimensional balls built up inductively, by dimension, as follows.
A finite CW-complex C 0 of dimension 0 is a finite collection of points, each
comprising an open set in C 0 . These points are called open 0-cells.
A finite CW-complex C n of dimension n is formed from a finite CW-complex
C n−1
of dimension n − 1 and a disjoint union of finitely many n-dimensional balls
B1 , . . . , Bk by identification as follows. For each Bi , there is a continuous function
fi : ∂Bi → C n−1 called an attaching map. Let X n denote the disjoint union of
the space C n−1 and the balls B1 , . . . , Bk . Let ∼ denote the equivalence relation
generated by requiring that each element x of each set ∂Bi be equivalent to its
image fi (x) under the attaching map fi . Then C n = X n / ∼. This space C n is
called the n-skeleton of C.
Note that the sets int(Bi ) are embedded in C n . They are called the open n-cells
of the complex C n .
Definition 16.18 (Simple subdivision). Suppose that C is a CW-complex and
that C is an open cell in C of dimension n. Then C is the interior of an n-dimensional
ball B that has been attached to the n−1 dimensional skeleton C n−1 by an attaching
map f : ∂B → C n−1 . Let B denote an (n−1)-dimensional ball in B with ∂B ⊂ ∂B
and with int(B ) ⊂ int(B) that divides B into two n-dimensional balls B1 and B2 .
Assume in addition that f |∂B : ∂B → C n−1 actually takes ∂B into C n−2 . Then
we may form a new CW-complex by omitting the open n-cell int(B), attaching one
new (n − 1) cell by the attaching map f |∂B , and attaching two new n-cells by the
attaching maps f ∪ id|∂B1 and f ∪ id|∂B2. We call the resulting CW-complex a
simple subdivision of C.
Corollary 16.19. The Euler characteristic χ(C) of a finite CW-complex is
invariant under simple subdivision.
There is a stronger invariance theorem that we shall assume but not prove.
Theorem 16.20 (Topological invariance of Euler characteristic). If C and D are
two finite CW-complexes whose underlying identification spaces are homeomorphic,
then χ(C) = χ(D). In fact, if the homology of C is isomorphic to the homology
16.2. EULER CHARACTERISTIC 145
of D, then χ(C) = χ(D), which happens for example when C and D are homotopy
equivalent.
Remark. The theorem on topological invariance of Euler characteristic allows
us to speak of the Euler characteristic of a sphere or ball, of a compact 2-manifold,
of n-space, etc. End remark.
Corollary 16.21. The Euler characteristic of the n-sphere Sn is equal to 2
in even dimensions and is equal to 0 in odd dimensions.
Proof. The n-sphere can be realized as a CW-complex with one 0-cell and
one n-dimensional cell. If n is even, then χ(Sn ) = 10 + 1n = 2. If n is odd, then
χ(Sn ) = 10 − 1n = 0.
But χ(Mj ) = χ(Mj ) − 1 since Mj has one fewer 2-cells than does Mj . The intersec-
tion of M1 and M2 after identification of boundaries is a 1-sphere, which has Euler
characteristic 0. The corollary follows.
Theorem 16.26 (Euler characteristic of a 2-manifold). Suppose that the com-
pact, connected 2-manifold M is the connected sum of S2 with k copies of the torus
and = 0, 1, or 2 copies of the projective plane. Then the Euler characteristic of
M is
χ(M ) = 2 + (k · 0 − 2k) + ( · 1 − 2) = 2 − 2k − .
Proof. The Euler characteristic of S2 is 2. Each of the k-connected summands
that is a torus adds the characteristic 0 of the torus and subtracts 2. Each of the
-connected summands that is a projective plane adds the characteristic 1 of the
projective plane and subtracts 2.
Corollary 16.27. Two compact connected 2-manifolds M1 and M2 are home-
omorphic iff the following two conditions are satisfied:
(1) Both are orientable or both are nonorientable.
(2) χ(M1 ) = χ(M2 ).
Proof. The two conditions are clearly necessary in order that M1 and M2
be homeomorphic. Suppose both are satisfied. We may realize Mj as a connected
sum of the 2-sphere with kj copies of T2 and j = 0, 1, or 2 copies of P2 . If both
are orientable, then 1 = 2 = 0. If both are nonorientable, then 1 = 2 = 2 if
χ(M1 ) = χ(M2 ) is even, and 1 = 2 = 1 if χ(M1 ) = χ(M2 ) is odd. Hence, in any
case, 1 = 2 . It follows that k1 = k2 . Hence M1 and M2 are homeomorphic.
Theorem 16.28 (Euler characteristic of an odd-dimensional manifold). The
Euler characteristic of an odd-dimensional manifold M is 0.
Proof. Though the result is true in general by homological considerations
(Poincaré duality), we shall only prove it for manifolds that are triangulated. We
employ Alexander’s algebraic representation which describes each simplex as the
product (join) of its vertices and the manifold itself as the formal sum M of its
principal simplexes (simplexes of maximal dimension). If σ is a simplex of M , then
M can be factored in the form M = σL + K. The subcomplex L is called the
link of σ in M , and the product σL is called the star of σ in M . In the standard
(combinatorial) triangulations of an n-manifold, the link of a k-simplex is always
a sphere of dimension n − k − 1. We assume this to be true of our triangulations.
The algebraic product of disjoint simplexes is called their join.
We will sum the Euler characteristics of the links of every simplex of M , di-
mension by dimension. We calculate the answers in two different ways. In the
first, we use our knowledge that vertex links are (n − 1)-spheres, edge links are
(n − 2)-spheres, triangle links are (n − 3)-spheres, etc. In the second, we consider
the contribution of each simplex to all link Euler characteristics and sum over all
simplexes.
First calculation: The Euler characteristic of a link is 2 for simplexes of even
dimension, 0 for simplexes of odd dimension. Therefore the sum over all simplexes
is
2 · #(0) + #(2) + #(4) + · · · + #(n − 1) ,
where #(k) denotes the number of simplexes of dimension k.
16.2. EULER CHARACTERISTIC 147
2 3 4 n+1
χ(link(v)) = #(1) − #(2) + #(3) − · · · + #(n) ,
v a vertex
1 1 1 1
3 4 n+1
χ(link(e)) = +#(2) − #(3) + · · · − #(n) ,
2 2 2
e an edge
4 n+1
χ(link(t)) = +#(3) − · · · + #(n) ,
3 3
t a triangle
... ...
n+1
χ(link(α)) = +#(n) .
n
dim(α)=n−1
The link of an n-simplex is empty and thus does not appear. Summing, we
obtain
2 · #(1) + #(3) + #(5) + · · · + #(n) .
Equating our two calculations, we find
2 · #(0) + #(2) + #(4) + · · · + #(n − 1) = 2 · #(1) + #(3) + #(5) + · · · + #(n) ,
or χ(M ) = 0.
Knot
Seifert Circles
In the example, circles = 3 and crossings = 6. Therefore the Euler characteris-
tic is 3 − 6 = −3. If we were to fill in the boundary of the surface with a disk, we
would increase the Euler characteristic by 1 to form a closed orientable surface of
Euler characteristic = −2. Thus the Seifert surface is a disk with two handles.
16.3. Exercises
16.2. Solve Exercise 16.14 on page 142.
16.2. Solve Exercise 16.2 on page 143.
16.3. Solve Exercise 16.16 on page 144.
16.4.Solve Exercise 16.24 on page 145.
16.5. Construct orientable Seifert surfaces for the two knots in Figure 7. Cal-
culate the Euler characteristic of each surface.
16.6. Given the edge pairings in Figures 8 and 9, determine whether the re-
sulting surfaces are orientable and calculate the Euler characteristic.
150 16. ORIENTATION AND EULER CHARACTERISTIC
e−1
c−1
d−1
a−1
a
b−1
c
Figure 8. An edge-pairing
g c
d
g −1
b
e
f −1
a−1
a
j
i
i
e
h−1
c−1
b
h
d f
j
17.1. Setting
We first have to prepare the setting.
We consider functions f (z) of one complex variable z which are rational in the
sense that they are the quotient f (z) = p(z)/q(z) of two polynomials p(z) and q(z)
with complex coefficients.
The 2-dimensional sphere S2 can be interpreted as the Riemann sphere, which is
the union of the complex plane C with a single point at infinity. The correspondence
which supports this interpretation is stereographic projection, which we described
in detail in the section on Pythagorean triples, Volume 1, Chapter 2.
A rational function of one complex variable can therefore be interpreted as a
function that maps the two-dimensional sphere S2 onto itself. The poles of the
rational function are the points that are mapped to infinity. The image of infinity
can be calculated by taking the limit of the function as domain points approach
infinity.
Such a function can be iterated to form a discrete dynamical system, and the
properties of such systems are unendingly fascinating: Near which points is the
system infinitely complicated (the Julia set)? Near which points is the system
simple (the Fatou set)? The computer visualizations of such sets are intricate and
beautiful. The study of even the family of quadratic polynomials of the form z 2 + c
is complicated and leads to the famous Mandelbrot set, which is parametrized by
the complex numbers c. A first approximation to the complexity is supplied by the
singularities of the rational map, and the nature of the singularities is controlled to
some extent by Euler characteristic as appearing in the Riemann-Hurwitz Theorem.
Before explaining that theorem, we need to describe the local properties of a rational
map.
p (t)
(−s(t), c(t))
There is obviously only one vector p (t) satisfying these conditions, namely
p (t) = (−s(t), c(t)). Since c(t) = cos(t), and s(t) = sin(t), the proof is complete.
(sin α) · r
(cos α) · r
α
(C, S)
cos β · sin α
1 β
α sin β · cos α
cos β · cos α
Proof. Graph the point (c, s) = (cos(α + β), sin(α + β)), and apply the pro-
jection principle to evaluate the four lengths indicated by arrows (Figure 3).
154 17. THE RIEMANN-HURWITZ THEOREM
It is obvious from the projection principle as applied to the diagram, that the
products that appear in the sum formulas have clear geometric meaning so that
cos(α + β) = c = cos(β) · cos(α) − sin(β) · sin(α)
and that
sin(α + β) = s = sin(β) · cos(α) + cos(β) · sin(α),
as required by the theorem.
Proof.
[r, θ]pol · [s, η]pol = r(cos θ + i sin θ) · s(cos η + i sin η)
= rs (cos θ · cos η − sin θ · sin η) + i(sin θ · cos η + cos θ · sin η)
= [rs, θ + η]pol .
vN ◦ f ◦ u−1
M (z) = z
n(p)
(3) If S1 = S = S2 , then
(deg(f ) − 1) · χ(S) = {n(p) − 1 | p ∈ S} ≥ 0.
is deg(f ) = {n(pi ) | i = 1, . . . , k}, and, consequently, q has only k = deg(f ) −
{n(pi ) − 1 | i = 1, . . . , k} preimages. We thus obtain the following equation:
χ(S1 ) = V (T1 ) − E(T1 ) + F (T1 )
= deg(f )(V (T2 ) − E(T2 ) + F (F2 )) − {n(p) − 1 | p ∈ S1 }
deg(f )χ(S2 ) − {n(p) − 1 | p ∈ S1 }
(3) If S1 = S = S2 , then the equation
(deg(f ) − 1) · χ(S) = {n(p) − 1 | p ∈ S} ≥ 0
is an immediate consequence of (2).
Proof of the Corollary 17.9 to the Riemann-Hurwitz Theorem.
(1) If χ(S) < 0, then (deg(f ) − 1) · χ(S) is both ≤ 0 and ≥ 0, hence = 0. Thus
deg(f ) = 1 and f is a homeomorphism.
If χ(S) = 0 so that S is a torus, then the (deg(f ) − 1) · χ(S) = 0 so
(2)
that {n(p) − 1 | p ∈ S} = 0. Consequently, each n(p) is 1 and f is a local
homeomorphism, hence a covering map by (1) of the Riemann-Hurwitz Theorem.
S is the 2-sphere, then χ(S) = 2 so that (deg(f ) − 1) ·
(3) If χ(S) > 0 so that
χ(S) = 2 deg(f ) − 2 = {n(p) − 1 | p ∈ S}.
f (N )
N
q
D3
p D2 A
D
D1
17.7. Exercises
17.1. For what surfaces S1 and S2 is it possible to construct a branched map
f : S1 → S2 ?
17.2. The study of branched maps from the 2-sphere to itself has spawned a
huge literature. Look up the Mandelbrot set on the internet. Find other references
there. Find the relationship between branched maps and fractals.
Bibliography
Plain Fun
(top recommendations for easy, but rewarding, pleasure).
[1] Hardy, G. H., A Mathematician’s Apology, Cambridge University Press, 2004 (eighth print-
ing).
[2] Pólya, G., How to Solve It, Princeton Univerity Press, 2004.
[3] Körner, T. W., The Pleasures of Counting, Cambridge University Press, 1996.
- More Fun
[4] Davis, P. J. and Hersh, R., The Mathematical Experience, Houghton Mifflin Company, 1981.
[5] Rademacher, H., Higher Mathematics from an Elementary Point of View, Birkhäuser, 1983.
[6] Hilbert, D., and Cohn-Vossen, S., Geometry and the Imagination, (translated by P. Nemeyi),
Chelsea Publishing Company, New York, 1952. [College level exposition of rich ideas from
low-dimensional geometry, with many figures.]
[7] Dörrie, H., 100 Great Problems of Elementary Mathematics: Their History and Solution,
Dover Publications, Inc., 1965, pp. 108-112. [We learned our first proof of the fundamental
theorem of algebra here.]
[8] Courant, R. and Robbins, H., What is Mathematics?, Oxford University Press, 1941.
Classics
(a chance to see the thinking of the very best, in chronological order).
[9] Euclid, The Thirteen Books of Euclid’s Elements, Vol. 1-3, 2nd Ed., (edited by T. L. Heath)
Cambridge University Press, Cambridge, 1926. [Reprinted by Dover, New York, 1956.]
[10] Archimedes, The Works of Archimedes, edited by T. L. Heath, Dover Publications, In.,
Mineola, New York, 2002. See also the exposition in Pólya, G., Mathematics and Plausible
Reasoning, Vol. 1. Induction and Analogy in Mathematics, Chapter IX. Physical Mathemat-
ics, pp. 155-158, Princeton University Press, 1954. [How Archimedes discovered the integral
calculus.]
[11] Wallis, J., in A Source Book in Mathematics, 1200-1800, edited by D. J. Struik., Harvard
University Press, 1969, pp. 244-253. [Wallis’s product formula for π.]
[12] Gauss, K. F., General Investigations of Curved Surfaces of 1827 and 1825, Princeton Uni-
versity Library, 1902. [Available online. Difficult reading.]
[13] Fourier, J., The Analytical Theory of Heat, translated by Alexander Freeman, Cambridge
University Press, 1878. [Available online, 508 pages. The introduction explains Fourier’s
thoughts in approaching the problem of the mathematical treatment of heat. Chapter 3 ex-
plains his discovery of Fourier series.]
[14] Riemann, B., Collected Papers, edited by Roger Baker, Kendrick Press, Heber City, Utah,
2004. [English translation of Riemann’s wonderful papers.]
159
160 BIBLIOGRAPHY
[15] Poincaré, H., Science and Method, Dover Publications, Inc., 2003. [Discusses the role of the
subconscious in mathematical discovery.] Also, The Value of Science, translated by G. B.
Halstead, Dover Publications, Inc., 1958.
[16] Klein, F., Vorlesungen über Nicht-Euklidische Geometrie, Verlag von Julius Springer, Berlin,
1928. [In German. An algebraic development of non-Euclidean geometry with respect to the
Klein and projective models. Beautiful figures. Elegant exposition.]
[17] Hilbert, D., Gesammelte Abhandlungen (Collected Works), 3 volumes, Springer-Verlag,
1970. [In German. The transcendence of e and π appears in Volume 1, pp. 1-4. Hilbert’s
space-filling curve appears in Volume 3, pp. 1-2.]
[18] Einstein, A., The Meaning of Relativity, Princeton University Press, 1956.
[19] Thurston, W. P., Three-Dimensional Geometry and Topology, edited by Silvio Levy, Prince-
ton University Press, 1997. [An intuitive introduction to dimension 3 by the foremost ge-
ometer of our generation.]
[20] W. P. Thurston’s theorems on surface diffeomorphisms as exposited in Fathi, A., and Lau-
denbach, F., and Poénaru, V., Travaux de Thurston sur les Surfaces, Séminaire Orsay,
Société Mathématique de France, 1991/1979. [In French.]
History
(concentrating on famous mathematicians).
[21] Bell, E. T., Men of Mathematics, Simon and Schuster, Inc., 1937. [The book that convinced
me that mathematics is exciting and romantic.]
[22] Henrion, C., Women of Mathematics, Indiana University Press, 1997.
[23] Dunham, W., Journey Through Genius, Penguin Books, 1991.
Supporting Textbooks
- Topology
[24] Munkres, J. R., Topology, a First Course, Prentice-Hall, Inc., 1975. [The early chapters
explain the basics of topology that form the prerequisites for the latter half of this book. The
later chapters contain rather different views of some of the later theorems in our book.]
[25] Massey, W. S., Algebraic Topology: An Introduction. Springer-Verlag, New York -
Heidelberg-Berlin, 1967 (Sixth printing: 1984), Chapter I, pp. 1-54. [A particularly nice
introduction to covering spaces.]
[26] Hatcher, A., Algebraic Topology , Cambridge University Press, 2001. [A very nice introduc-
tion to algebraic topology, a bit of which we need in Volume 2.]
[27] Munkres, J. R., Elements of Algebraic Topology, Addison-Wesley, 1984. [Another nice in-
troduction.]
[28] Alexandroff, P., Elementary Concepts of Topology, translated by Alan E. Farley, Dover
Publications, Inc., 1932. [A wonderful small book.]
[29] Alexandrov, P. S., Combinatorial Topology, 3 volumes, translated by Horace Komm, Gray-
lock Press, Rochester, NY, 1956.
[30] Seifert, H., and Threlfall, W., A Textbook of Topology, translated by Michael A. Goldman;
and Seifert, H., Topology of 3-Dimensional Fibered Spaces, translated by Wolfgang Heil,
Academic Press, 1980. [Available online.]
[31] Hurewicz, W., and Wallman, H., Dimension Theory, Princeton University Press, 1941. [See
Chapters 4, 5, and 6 of Volume 2.]
BIBLIOGRAPHY 161
- Algebra
[32] Herstein, I. N., Abstract Algebra, third edition, John Wiley & Sons, Inc., 1999. [See our
Chapter 6 of Volume 1.]
[33] Dummit, D. S., and Foote, R. M., Abstract Algebra, third edition, John Wiley & Sons, Inc.,
2004. [See our Chapter 6 of Volume 1.]
[34] Hardy, G. H., and Wright, E. M., An Introduction to the Theory of Numbers, fourth
edition, Oxford University Press, 1960. [See our Chapter 5 of Volume 1.]
- Analysis
[35] Apostol, T. M., Mathematical Analysis , Addison-Wesley, 1957. [Good background for Rie-
mannian metrics in Chapter 1 of Volume 1, and also the chapters of Volume 3.]
[36] Lang, S., Real and Functional Analysis, third edition, Springer, 1993. [Chapter XIV gives the
differentiable version of the open mapping theorem. The proof uses the contraction mapping
principle. See our Chapter 12 of Volume 2 for the topological version of the open mapping
theorem.]
[37] Spivak, M., Calculus on Manifolds, W. A. Benjamin, Inc., New York, N. Y., 1965. [Good
background for Riemannian metrics in Chapter 1 of Volume 1 and for Volume 3.]
[38] Jänich, K., Vector Analysis, translated by Leslie Kay, Springer, 2001. [Good background for
Riemannian metrics in Chapter 1 of Volume 1 and for Volume 3.]
[39] Saks, S., Theory of the Integral, second revised edition, translated by L. C. Young, Dover
Publications, Inc., New York, 1964. [Wonderfully readable.]
[40] H. L. Royden, H. L., Real Analysis, third edition, Macmillan, 1988. [The place where we
first learned about nonmeasurable sets.]
[41] Flavors of Geometry, edited by Silvio Levy, Cambridge University Press, 1997.
[42] Alonso, J. M., Brady, T., Cooper, D., Ferlini, V., Lustig, M., Mihalik, M., Shapiro, M.,
Short, H., Notes on word hyperbolic groups, Group Theory from a Geometrical Viewpoint:
21 March — 6 April 1990, ICTP, Trieste, Italy, (E. Ghys, A. Haefliger, and A. Verjovsky,
eds.), World Scientific, Singapore, 1991, pp. 3–63.
[43] Benedetti, R., and Petronio, C., Lectures on Hyperbolic Geometry, Universitext, Springer-
Verlag, Berlin, 1992. [Expounds many of the facts about hyperbolic geometry outlined in
Thurston’s influential notes.]
[44] Bolyai, W., and Bolyai, J., Geometrische Untersuchungen, B. G. Teubner, Leipzig and
Berlin, 1913. (reprinted by Johnson Reprint Corp., New York and London, 1972) [Historical
and biographical materials.]
[45] Cannon, J. W., The combinatorial structure of cocompact discrete hyperbolic groups, Geom.
Dedicata 16 (1984), 123–148.
[46] Cannon, J. W., The theory of negatively curved spaces and groups, Ergodic Theory, Symbolic
Dynamics and Hyperbolic Spaces, (T. Bedford, M. Keane, and C. Series, eds.) Oxford
University Press, Oxford and New York, 1991, pp. 315–369.
[47] Cannon, J. W., The combinatorial Riemann mapping theorem, Acta Mathematica 173
(1994), 155–234.
[48] Cannon, J. W., Floyd, W. J., Parry, W. R., Squaring rectangles: the finite Riemann mapping
theorem, The Mathematical Heritage of Wilhelm Magnus — Groups, Geometry & Special
162 BIBLIOGRAPHY
[71] Newman, M. H. A., Elements of the Topology of Plane Sets of Points, Cambridge University
Press, 1939.[A good alternative introduction to the topology of the plane.]
- Volume 1, Chapter 1
[72] Feynman, R., The Character of Physical Law, The M.I.T. Press, 1989, p. 47.[All of
Feynman’s writing is fun and thought provoking.]
- Volume 1, Chapter 2
[73] Gilbert, W. J., and Vanstone, S. A., An Introduction to Mathematical Thinking, Pearson
Prentice Hall, 2005. [The place where I learned the algorithmic calculations about the
Euclidean algorithm. See our Chapter 2.]
- Volume 1, Chapter 3
- Volume 1, Chapter 4
[74] Reid, C., Hilbert, Springer Verlag, 1970. [A wonderful biography of Hilbert, with an extended
discussion of the Hilbert address in which he stated the Hilbert problems. See our Chapter
4.]
- Volume 1, Chapter 5
[75] Apostol , T. M., Calculus , Volume 1, Blaisdell Publishing Company, New York, 1961. [The
place where I first learned areas by counting. See our Chapter 5.]
- Volume 1, Chapter 6
[76] Hilton, P., and Pedersen, J., Approximating any regular polygon by folding paper, Math.
Mag. 56 (1983), 141-155. [Method for approximating many angles algorithmically by paper-
folding.]
[77] Hilton, P., and Pedersen, J., Folding regular star polygons and number theory Math. Intel-
ligencer 7 (1985), 15-26. [More paper-folding.]
[78] Burkard Polster, Variations on a Theme in Paper Folding, Amer. Math. Monthly 111 (2004),
39-47. [More paper-folding approximations to angles. See Chapter 6 and the impossibility of
trisecting an angle.]
164 BIBLIOGRAPHY
- Volume 1, Chapter 7
[79] Wagon, S., The Banach-Tarski Paradox, Cambridge University Press, 1994.[A wonderful
exposition of the Hausdorff-Banach-Tarski paradox, without the emphasis on the graph of
the free group. See our Chapter 7.]
[80] Coxeter, H. S. M., and Moser, W. O., Generators and Relations for Discrete Groups, second
edition, Springer-Verlag, 1964. [The place where I learned that groups can be viewed as graphs
(the Cayley graph or the Dehn Gruppenbild). See our Chapter 7 where we use the graph of
the free group on two generators and Chapter 25 where we use graphs as approximations
to non Euclidean geometry.]
- Volume 2, Chapter 13
[81] Peano, G. , Sur une courbe, qui remplit toute une aire plane, Mathematische Annalen 36
(1), 1890, pp. 157-160. [The first space-filling curve, described algebraically. See our Chapter
12.]
[82] Peano, G., Selected works of Giuseppe Peano, edited by Kennedy, Hubert C., and translated.
With a biographical sketch and bibliography, Allen & Unwin, London, 1973.
[83] Hilbert, D., Über die stetige Abbildung einer Line auf ein Flächenstück, Mathematische
Annalen 38 (3), 1891, pp. 459-460. [Hilbert gave the first pictures of a space-filling curve.
See our Chapter 12.]
[84] G. Pólya, Über eine Peanosche Kurve, Bull. Acad. Sci. Cracovie, A, 1913, pp. 305-313.
[Pólya’s triangle-filling curve. See our Chapter 12.]
[85] Lax, P. D., The differentiability of Pólya’s function, Adv. Math., 10, 1973, pp. 456-464.
[Lax recommends the non-isosceles triangle in Pólya’s construction since it simplifies the
description of the path followed to the point represented by a binary expansion. See our
Chapter 12.]
- Volume 2, Chapter 6
[86] Mandelbrot, B., The Fractal Geometry of Nature, W. H . Freeman & Co, 1982. [Mandelbrot
suggests the use of Hausdorff dimension as a means of recognizing sets that are locally
complicated or chaotic. He defines these to be fractals. See our Chapter 13.]
[87] Falconer, K. J., The Geometry of Fractal Sets, Cambridge University Press, 1985. [See
reference [86] and our Chapter 13.]
[88] Devaney, R. L., Differential Equations, Dynamical Systems, and an Introduction to Chaos
with Morris Hirsch and Stephen Smale, 2nd edition, Academic Press, 2004; 3rd edition,
Academic Press, 2013. [See reference [84] and our Chapter 13.]
[89] Moore, R. L., Concerning upper semi-continuous collections of continua, Trans. Amer.
Math. Sc. 27 (1925), pp. 416-428. [Moore shows that his topological characterization of the
plane or 2-sphere allows him to prove his theorem about decompositions of the 2-sphere. See
our Volume 2, Chapters 8 and 11.]
BIBLIOGRAPHY 165
[90] Wilder, R. L., Topology of Manifolds, American Mathematical Society, 1949 . [Our proof of
the topological characterization of the sphere is primarily modelled on Wilder’s proof, with
what we consider to be conceptual simplifications. See our Chapter 8.]
- Volume 2, Chapter 13
[91] Radó, T., Über den Begriff der Riemannschen Fläche, Acts. Litt. Sci. Szeged 2 (1925), pp.
101-121. [The first proof that 2-manifolds can be triangulated. See our Chapter 20.]
- Volume 2, Chapter 14
[92] Andrews, Peter, The classification of surfaces, Amer. Math. Monthly 95 (1988), 861-867l
[93] Armstrong, M. A., Basic Topology, McGraw-Hill, London, 1979.
[94] Burgess, C. E., Classification of surfaces, Amer. Math. Monthly 92 (1985), 349-354.
[95] Francis, George K., Weeks, Jeffrey R., Conway’s ZIP proof, Amer. Math. Monthly 106
(1999), 393-399.
- Volume 2, Chapter 15
[96] Rolfsen, D., Knots and Links, AMS Chelsea, vol 346, 2003. [See our Chapter 22.]
For the entirety of Volume 3, see the references above taken from our article in
Flavors of Geometry, beginning with reference [41].
- Volume 3, Chapter 3
[97] Misner, C. W., and Thorne, K. S., and Wheeler, J. A., Gravitation, W. H. Freeman and
Company, 1973.
[98] Abelson, H., and diSessa, A., Turtle Geometry, MIT Press, 1986. [The authors use the paths
of a computer turtle to model straight paths on a curved surface.]
This is the second of a three volume collection devoted to the geometry,
topology, and curvature of 2-dimensional spaces. The collection provides
a guided tour through a wide range of topics by one of the twentieth
century’s masters of geometric topology. The books are accessible to
college and graduate students and provide perspective and insight to
mathematicians at all levels who are interested in geometry and topology.
The second volume deals with the topology of 2-dimensional spaces. The
attempts encountered in Volume 1 to understand length and area in the
plane lead to examples most easily described by the methods of topology
(fluid geometry): finite curves of infinite length, 1-dimensional curves of positive area, space-
filling curves (Peano curves), 0-dimensional subsets of the plane through which no straight
path can pass (Cantor sets), etc. Volume 2 describes such sets. All of the standard topological
results about 2-dimensional spaces are then proved, such as the Fundamental Theorem of
Algebra (two proofs), the No Retraction Theorem, the Brouwer Fixed Point Theorem, the
Jordan Curve Theorem, the Open Mapping Theorem, the Riemann-Hurwitz Theorem, and
the Classification Theorem for Compact 2-manifolds. Volume 2 also includes a number of
theorems usually assumed without proof since their proofs are not readily available, for example,
the Zippin Characterization Theorem for 2-dimensional spaces that are locally Euclidean, the
Schoenflies Theorem characterizing the disk, the Triangulation Theorem for 2-manifolds, and
the R. L. Moore’s Decomposition Theorem so useful in understanding fractal sets.
MBK/109