You are on page 1of 22

Mathematics 4530

Tychonoff s Theorem
Ken Brown, Cornell University, October 2013
1. Introduction
Tychonoffs theorem asserts that the product of an arbitrary family of compact
spaces is compact. This is proved in Chapter 5 of Munkres, but his proof is not
very straightforward. The proof Ill give below follows a paper by David Wright
(Proceedings of the American Mathematical Society 120 (1994), 985987). As a
warmup, lets start with two factors.
2. The Baby Tychonoff Theorem
Theorem. If X and Y are compact, then so is X Y .
Note that this immediately extends to arbitrary finite products by induction on
the number of factors. This yields:
Corollary. A subset of Rn is compact if and only if it closed and bounded.
Proof of the corollary. If X Rn is compact, then it is closed and bounded by the
same proof we used in class when n = 1. Conversely, if X is closed and bounded,
then X is a closed subset of a rectangle R = I1 In , where each Ii is a closed
interval in R. Since each Ii is compact, so is the product R; the closed subset X
of R is therefore also compact.

To prove a space X is compact, one usually proves that if U is a family of open
sets that covers X, then a finite subcollection covers X. Sometimes, however, it is
more convenient to prove the contrapositive: If U is a family of open sets such that
no finite subcollection covers X, then U does not cover X. This is what we will do
below.
Proof of the theorem. Let W be a collection of open subsets of X Y such that no
finite subcollection covers X Y ; we will show that W does not cover X Y .
Claim 1. There exists x0 X such that no open tube U Y with x0 U is finitely
covered by W.
Proof. If this is false, then every x X has a neighborhood Ux such that Ux Y is
finitely covered. By compactness of X, finitely many of these sets Ux cover X, so
finitely many of the tubes Ux Y cover X Y . This contradicts the assumption
that X Y is not finitely covered.
Claim 2. There exists y0 Y such that no open rectangle U V containing (x0 , y0 )
is finitely covered by W.
Proof. If this is false, then for every y Y there is a finitely covered open rectangle
Uy Vy containing
of Y , there is a finite subset F Y
S (x0 , y). By compactness
T
such that Y = yF Vy . Set U := yF Uy . Then U is a neighborhood of x0 , and
the tube
[
[
U Y =
U Vy
Uy Vy
yF

yF

is finitely covered, contradicting Claim 1.


1

Claim 2 immediately implies the theorem. Indeed, we have a point z := (x0 , y0 )


such that no basic open set containing z is finitely covered by W. In particular,
no
S
basic open set containing z can be contained in a set W W, so z
/ W W W .
This shows that W does not cover X Y .

Remark. We did not have to use the contrapositive of the condition for compactness.
In fact, it is more straightforward to just use the original definition, and to base the
proof on the tube lemma; see Munkres, pp. 167168. (The tube lemma asserts that
in a product X Y with Y compact, any neighborhood of a slice x Y contains
a tube U Y , where U is a neighborhood of x; the proof of this is very similar to
the proof of Claim 2.) The advantage of our convoluted proof, however, is that it
extends easily to infinitely many factors.
3. Countably Many Factors
In preparation for treating infinitely many factors, we record a slight generalization of the argument used in the proof of Claim 2 above.
Lemma. Let W be a family of open sets in a product X Y Z. Assume there is a
point x0 X such that no open set U Y Z with x0 U is finitely covered by W.
If Y is compact, then there is a point y0 Y such that no open set U V Z with
(x0 , y0 ) U V is finitely covered by W.
Proof. Suppose no such y0 exists. Then for every y Y there is a finitely covered
open set Uy Vy Z with x0 Uy and
of Y , there is
S y Vy . By compactness
T
a finite subset F Y such that Y = yF Vy . Set U := yF Uy . Then U is a
neighborhood of x0 , and the set
[
[
U Y Z =
U Vy Z
Uy Vy Z
yF

is finitely covered, contradicting the hypothesis.

yF

We can now easily generalize the baby argument to treat countably many
factors:
Q
Theorem. Let X = i=1 Xi , where each Xi is compact. Then X is compact.
Proof. Let W be a family of open sets that does not finitely cover X; we will
construct a point x = (x1 , x2 , . . . ) such that no neighborhood of x is finitely covered.
Note first that there is a point x1 X1 such that no open tube U X2 X3
with x1 U is finitely covered. The proof of this assertion is the same as the proof
of Claim 1 above, with X2 X3 playing the role of Y . Next, we can find
x2 X2 such that no open rectangle U V X3 X4 with (x1 , x2 ) U V is
finitely covered. This follows from the lemma if we view X as X1 X2 (X3 ).
Continuing in this way, we inductively define x1 , x2 , x3 , . . . such that for each n, no
basic open set of the form U1 Un Xn+1 , with xi Ui for i n, is finitely
covered. For the inductive step, view X as (X1 Xn1 ) Xn (Xn+1 )
and apply the lemma.
We now have a point x = (x1 , x2 ,S
. . . ) X such that no basic neighborhood of x
is finitely covered by W. Thus x
/ W W W .


4. Grown-up Tychonoff
Finally, here is the full-fledged Tychonoff theorem:
Theorem.
Given an arbitrary family (X )J of compact spaces, their product
Q
X := J X is compact.
Proof. Once again, we let W be a family of open sets that does not finitely cover X,
and we construct x = (x )J such that no neighborhood of x is finitely covered.
We may assume by the well-ordering theorem that the index set J is well-ordered,
and we construct x by transfinite induction so that no basic open set of the form
Y
Y
U
X ,

>

with x U for all , is finitely covered. For the inductive step, assume that
x has been defined
with the desired
 for all < , and apply the lemma,
Q
Q property
viewing X as
X

<
> . The details are essentially the same
as in the countable case.

Remark. The proof used the axiom of choice, in the form of the well-ordering theorem. This is unavoidable. Indeed, logicians have shown that Tychonoffs theorem
cannot be proved without the axiom of choice.
5. Application: Invariant Means
Having gone to all the trouble of proving the well-ordering theorem and the
general form of Tychonoffs theorem, we now give a (somewhat weird) application
of it. As is typical of results that require the axiom of choice, it proves the existence
of something that one could never hope to actually construct concretely. We begin
with a nave question:
Is there a sensible way of associating an average value to every bounded, doubly
infinite sequence of real numbers, such as
(1)

. . . , 0, 1, 0, 1, 0, 1, . . .?

To phrase the question precisely, consider sequences a = (an )nZ , i.e., elements of
RZ . We denote by B the set of all such sequences that are bounded. It is a real
vector space. We will not make use of any topology on B, but we recall in passing
that there is a standard one, different from the product topology, that is useful in
analysis: One puts a norm on B by setting kak = supnZ |an |, and this yields a
metric d(a, b) := ka bk and hence a topology.
A mean on B is a linear map : B R such that
inf an (a) sup an
n

for all a B. Means exist in great abundance. For example, we can take a finite
subset F Z and define a mean by averaging over F :
1 X
(a) :=
an .
|F |
nF
P
More generally, we can assign a weight wn 0 to each n Z, with n wn = 1,
and use the weighted average
X
(a) :=
an wn .
n

The shift operator a 7 as on B is defined by (as )n = an+1 . A mean is said to be


invariant, or shift-invariant, if (as ) = (a) for all a B. It is reasonable to expect
a sensible averaging procedure to be shift-invariant; this just says that we can talk
about the average of a sequence like the one in (1) above, without having to choose
arbitrarily which element is considered to be in postion 0. But no one knows how
to construct an invariant mean. (In particular, the weighted averages above are
not invariant). The best we can do is to construct approximately invariant means
and then deduce, in a nonconstructive way, that invariant means exist. Here is the
constructive part:
Lemma. There is a sequence (n )n1 of means such that for all a B,
lim |n (as ) n (a)| = 0.

Proof. Let n (a) := (1/n)

Pn

i=1

ai . Then

|n (as ) n (a)| = |an+1 a1 |/n 2kak/n,


which tends to 0 as n .

There are several known methods for doing the nonconstructive part of the proof
that invariant means exist. The one we will use is to topologize the set M of all
means in such a way that it becomes a compact space (via Tychonoff). It will then
follow that the approximately invariant sequence n has a cluster point , which is
invariant. Here are the details.
View M as a subset of RB by identifying a mean with the indexed family
((a))aB . We then give RB the product topology and M the subspace topology.
Thus our topology is cooked up to make the evaluation map 7 (a) a continuous
map M R for each a B. Note that the definition of mean actually makes
M a subset of a product of closed intervals:
Y
M
[m(a), M (a)],
aB

where m(a) = inf n an and M (a) = supn an .


Proposition. M is compact.
Q
Proof. We know that X := aB [m(a), M (a)] is compact, so it suffices to show
that M is a closed subset of X. Now an element X is a mean if and only if it
satisfies
(a + b) (a) (b) = 0
for all a, b B and
(a) (a) = 0
for all a B and R. This exhibits M as an intersection of sets of the form
F 1 (0) for various continuous maps F : X R, so M is closed.

We now appeal to the BolzanoWeierstrass property of compact spaces:
Proposition. If X is a compact space, then every sequence (xn )n1 in X has a
cluster point, i.e., there is a point x X such that every neighborhood of x contains
xn for infinitely many n.

Proof. For each x that is not a cluster point, there is a neighborhood Ux of x that
contains xn for only finitely many n. Clearly no finite subcollection
of the Ux can
S
cover X, so the family of all Ux does not cover X. Since x Ux contains all points
that are not cluster points, it follows that there must exist at least one cluster
point.

We can now prove our promised existence theorem:
Theorem. There exists an invariant mean : B R.
Proof. Take a sequence (n ) as in the lemma, and let be a cluster point. To show
that is invariant, let a B be arbitrary. We will argue that for a suitable large n,
(2)

(as ) n (as ) n (a) (a).

More precisely, fix  > 0 and let





U := M |(a) (a)| <  and |(as ) (as )| <  .
Then U is a neighborhood of in M. [In fact, it is a basic neighborhood, defined
by imposing conditions on two coordinates of a general .] We therefore have
n U for infinitely many n. Take such an n large enough that |n (as )n (a)| < .
Then the three approximations in (2) are all valid, with error less than , whence
|(as ) (a)| < 3. Since  is arbitrary, it follows that (as ) = (a).


Mathematics 4530

Prelim 1

October 22, 2013

Do any 4 of the following 6 problems, and show clearly which 4 you have chosen. You
may do additional problems for fun and a tiny amount of extra credit. Your proofs
may use any result proved in class, in the assigned reading, or in the homework,
unless you are explicitly asked to prove such a result.
1. Recall that a metric space X is said to be bounded if there is a number M such
that d(x, y) M for all x, y X. If X is a finite union of balls, show that X is
bounded.
2. Define each of the following terms; verbosity is not necessary.
(a) topology on a set X
(b) topological space
(c) subspace topology
(d) product topology (on an arbitrary product, possibly infinite)
(e) connected topological space
(f) compact topological space
3. Let X be a topological space. Given B A X, topologize A as a subspace
of X, and then topologize B as a subspace of A. Show that the resulting topology
on B is the same as the topology that B inherits as a subspace of X.
4. Let {A }J be a family of subsets of a metric space X. Define f : X R by
f (x) := d(x, A ),
and let f : X RJ be the map with components f , where RJ :=
the product topology. Is f continuous? Justify your answer.

R is given

5. Give a proof or counterexample for each of the following statements.


(a) If A and B are connected subspaces of a topological space and A B =
6 , then
A B is connected.
(b) If A and B are connected subspaces of a topological space, then A B is
connected.
6. Give a proof or counterexample for each of the following statements.
(a) If A and B are compact subspaces of a topological space, then AB is compact.
(b) If A and B are compact subspaces of a Hausdorff space, then A B is compact.

Brief solutions to Prelim 1


1. A ball B of radius r is bounded because d(x, y) 2r for all x, y B by the
triangle inequality. So it suffices to show that a finite union of bounded sets is
bounded. A straightforward induction reduces this to the case of two bounded sets
A1 , A2 . Let M1 and M2 be bounds for A1 and A2 , respectively. We may assume
that A1 and A2 are both nonempty, in which case we can choose points ai Ai . I
claim that M := M1 + d(a1 , a2 ) + M2 is a bound for A1 A2 : Given x, y A1 A2 ,
if they are both in A1 , then d(x, y) M1 M , so the claim holds in this case.
Similarly, the claim holds if they are both in A2 . Finally, if they are in different
sets, say x A1 and y A2 , then two applications of the triangle inequality give
d(x, y) d(x, a1 ) + d(a1 , y)
d(x, a1 ) + d(a1 , a2 ) + d(a2 , y)
M1 + d(a1 , a2 ) + M2
= M.
2. See Munkres.
3. This was a homework problem, but heres the solution: The given topology on
B has {(U A) B} as its open sets, where U ranges over the open sets of X.
Since (U A) B = U (A B) = U B, this is the same as the subspace topology
that B inherits from X.
4. Yes, f is continuous because each component function is continuous.
5. (a) This is true; see Munkres, Theorem 23.3.
(b) This is false. A counterexample is provided by the circle, which is the union of
two closed arcs intersecting in two points.
6. Recall that compactness for a subspace A of a space X can be formulated as
follows (see Munkres, Lemma 26.1): If U is a family of open subsets of X that
covers A, then a finite subset of U covers A.
(a) This is true. If U is a family of open subsets of X that covers A B, then
finitely many cover A and finitely many cover B; hence finitely many cover
A B.
(b) This is also true. A and B are closed in X because X is Hausdorff, so A B
is closed. Therefore A B is compact, being a closed subset of the compact
space A.

Mathematics 4530
Prelim 2, due December 5, 2013
What follows is a generalization of the calculation of 1 (S 1 ). It is essentially a
handout, but with all the proofs removed. Your task is to fill in the proofs in
Sections 16. [You may read Section 7 for fun if you want, but there is nothing
in it for you to write up.] I have given a clear indication of the places where you
are expected to fill in a proof. I have also labeled some proofs as extra credit to
keep the exam from being too long; they are not necessarily more difficult than the
other proofs.
You may use any result proved in class, in the assigned reading, or in your
homework. You may not use any sources other than your textbook and class notes.
You may not use results from unassigned homework problems or from sections of
the book that you have not been told to read. Andrew and I will be glad to clear
up any ambiguities; you may not discuss the exam with anyone else.
1. Introduction
A crucial feature of the exponential covering p : R S 1 is that two points in the
same fiber differ by an integral multiple of 2. This was used in the definition of the
degree homomorphism 1 (S 1 ) Z, which was then proved to be an isomorphism.
We wish to generalize this to other covering maps, the so-called regular ones. As a
consequence, one can calculate many fundamental groups with no more effort than
was required for S 1 .
2. Group actions
We have defined a group to be a set G together with a binary operation
satisfying certain axioms. This may give you the wrong impression of group theory,
in that it gives you no clue as to how groups arise in nature. The way groups arise
in nature is that they act on things, thereby exhibiting the symmetry that those
things have.
Definition 1. Let G be a group and X a topological space. By an action of G
on X we mean a function G X X, denoted (g, x) 7 g x, satisfying:
(1) g (h x) = (g h) x for all g, h G and x X.
(2) 1 x = x for all x X, where 1 is the identity element of G.
(3) For each g G, the map x 7 g x is continuous.
It follows that the map x 7 g x, which we will call the action of g, is actually a
homeomorphism. Its inverse is given by the action of g 1 . Group actions are studied
extensively in algebra courses, such as Math 4340. For our purposes, however, it
suffices to have some examples.
Example 1. The additive group Z acts on R by (n, x) 7 2n + x for n Z and
x R. (We say Z acts on R by translation.) This is the action thats implicit in
our work on the exponential covering map, which led to the calculation 1 (S 1 )
= Z.
Example 2. The multiplicative group {1} acts on S 2 by scalar multiplication.
This action is related to the double covering S 2 P 2 mentioned in class, where
P 2 is the projective plane.
1

Example 3. More generally, whenever we have a 2-fold covering map p : X Y ,


there is an associated action of a group of order 2 on X. This is essentially what
you proved in additional problem 6 on Assignment 10.
Example 4. Let Zn be the additive group of integers mod n. There is an action of
Zn on S 1 , in which the integer k mod n acts as rotation through 2k/n radians.
Alternatively, if we replace Zn by the isomorphic group consisting of the nth roots
of unity in C, then we can describe the action using multiplication of complex
numbers.
Example 5. The concept of Cayley graph, which I will explain in Section 7, leads
to many examples of group actions. For example, the snowflake that covers the
figure 8 is the Cayley graph of the free group F2 ; as a result, there is an action of
F2 on the snowflake.
To simplify the notation in what follows, we will always use multiplicative notation for our groups (i.e., we will write gh instead of g h), and similarly we will
write (g, x) 7 gx for group actions.
Every one of the examples above has the property that it is fixed-point free, i.e.,
the action of any g 6= 1 has no points x such that gx = x. The significance of this
for us is that if two points x, x0 are in the same G-orbit, i.e., if there is a g G
such that gx = x0 , then the element g is unique. [If gx = x0 and hx = x0 , then
gx = hx = h1 gx = x = h1 g = 1 = g = h.] We can think of g as
something like the difference between x and x0 . For the case of Z acting on R as
in Example 1, it really is the difference (divided by 2).
3. Regular G-covers
Definition 2. Let G be a group. A regular G-cover is a covering map p : X Y
together with a fixed-point-free action of G on X whose orbits are the fibers of p.
Lets spell this out explicitly:
Lemma 1. Let p : X Y be a covering map with an action of a group G on X.
Then p is a regular G-cover if and only if it satisfies the following two conditions:
(a) p(gx) = p(x) for all g G, x X.
(b) If p(x) = p(x0 ) then there is a (unique) g G such that x0 = gx.
Proof. *** Your proof goes here. ***

Remark 1. The group action is useful in connection with path liftings. Namely, if
is a path in Y and
is the lift starting at some point in the fiber over (0), then
we can construct all the other liftings of by applying the group action.
*** Your explanation goes here. ***
We have already seen many examples of regular G-covers. Indeed, every example
in the previous section has an associated regular G-cover, which was mentioned
explicitly except in Example 4. [You can ask me if you cant figure out the cover
that goes with that example.] On the other hand, there are many covering maps
that are not regular because they do not have enough symmetry. Indeed, conditions
(a) and (b) imply, intuitively, that a regular G-cover looks the same no matter where
we stand in a given fiber. For example:

Lemma 2. Let p : X Y be a regular G-cover for some group G, and let be a


closed curve in Y that admits a closed lift
. Then every lift of is closed.
Proof. *** Your proof goes here. ***

One can deduce from this that not all covering maps are regular.
Example 6. *** Your example goes here. ***
[To make it interesting, make sure that X and Y are both connected.]

4. The generalized degree homomorphism


Recall that we defined the degree of a closed curve in S 1 by lifting it to R and
measuring the difference between the starting and ending points of the lift. We can
do something similar for any regular cover.
Let p : X Y be a regular G-cover. Pick a basepoint y Y and a basepoint
x p1 (y). Given a path homotopy class [] 1 (Y, y), let
be the lift of
starting at x. Then
ends in the fiber p1 (y), so there is a unique g G such that

(1) = gx. We set


([]) := g.
In case p is the exponential covering of S 1 , this is precisely the degree defined in
class. Imitating the four steps that occurred in our calculation of 1 (S 1 ), we will
prove:
Theorem 1. Suppose that X is simply connected. Then
: 1 (Y, y) G
is an isomorphism.
The four steps are given in the following four lemmas, valid for any regular
G-covering.
Lemma 3. is well defined.
Proof. *** Your proof goes here. ***

Lemma 4. is a group homomorphism.


Proof. *** Your proof goes here. ***

Lemma 5. If X is path connected, then is surjective.


Proof. *** Your proof goes here. ***

Lemma 6. If X is simply connected, then is bijective.


Proof. *** Your proof goes here. ***

5. Examples
Example 7. Combining Theorem 1 and Example 2, we obtain:
*** Your conclusion goes here, with full justification. ***
Example 8. Applying Theorem 1 to a suitable covering of the torus, we obtain:
*** Your conclusion goes here, with full justification. ***
Example 9. I promised that I would prove that the fundamental group of the figure 8
is the free group F2 on two generators. One way to do this is to apply Theorem 1 to
the snowflake cover of the figure 8. I will sketch how this can be done in Section 7.
A different proof will be given in class.
6. Existence of regular G-covers
It is natural to ask how widely applicable Theorem 1 is. In this section we will
see that, in principle, Theorem 1 can be used to calculate the fundamental group
of every reasonable space. Here reasonable has a somewhat technical definition:
Definition 3. A space Y is said to be semilocally simply connected (SLSC) if
every point y Y has a neighborhood U such that the inclusion U , Y induces
the trivial homomorphism 1 (U, y) 1 (Y, y).
This condition may seem strange; but in fact, most spaces that one meets in
nature satisfy the even stronger condition of being locally contractible. [Feel free
to ask for examples if you have trouble seeing this.] Heres more evidence that the
SLSC condition is reasonable:.
Proposition 1. If Y admits a simply connected covering space, then Y is SLSC.
Proof. *** Your proof goes here. ***

Our goal in this section is to prove the following theorem:


Theorem 2. Suppose Y is path connected, locally path connected, and SLSC. Then
there exists a group G and a regular G-cover p : X Y such that X is simply
connected.
[Necessarily, G will be isomorphic to 1 (Y ).]
The first step is to figure out what X has to look like.
Lemma 7. (a) If U satisfies the condition in Definition 3, then every smaller
neighborhood of y also satisfies the condition.
(b) Suppose Y is SLSC and locally path connected. Let B be the collection of pathconnected open sets U such that 1 (U ) 1 (Y ) is the trivial homomorphism
for some (or every) basepoint in U . Then B is a basis for Y .
(c) For any U B let 1 , 2 be paths in U with 1 (0) = 2 (0) and 1 (1) = 2 (1).
Then [1 ] = [2 ] in the fundamental groupoid of Y .
Proof. *** Extra credit. ***

Let Y be a path-connected space with basepoint y0 , let G be its fundamental


groupoid, and let G0 G be the subset consisting of path classes starting at y0 .
Thus G0 is a set whose elements are equivalence classes a = [], where is a path
in Y with (0) = y0 . Let p : X Y be a covering map, and let x0 p1 (y0 ) be a

basepoint. Then one can use path lifting to define a function h : G0 X. Namely,
given [] G0 , we lift to a path
in X starting at x0 and set
h([]) =
(1).
Lemma 8. (a) h is well defined.
(b) h is surjective if X is path connected and bijective if X is simply connected.
Proof. *** Your proof goes here. ***

From now on we assume that Y is path connected, locally path connected, and
SLSC. We have just seen that if Y admits a simply connected covering space X,
then the points of X must be in 11 correspondence with the elements of the set G0 .
With this as motivation, we will prove Theorem 2 by starting with G0 and putting
a suitable topology on it. Let B be the basis for Y given in Lemma 7(b).
ea G0 be
Let a = [] be an element of G0 . For any U B with (1) U , let U
the set of path classes of the form a b, where b = [] for some path in U with
ea a basic neighborhood of a, and we will call a subset
(0) = (1). We will call U
of G0 open if it contains a basic neighborhood of each of its points.
ea as
Lemma 9. The open sets just defined form a topology on G0 , with the sets U
a basis.
Proof. *** Extra credit. ***

ea , then U
ea = U
ea .]
[Hint: It might help to show that if a1 U
1
As an aid to the intuition and a reminder of what were trying to do, we will
denote by X the set G0 with the topology that we have just defined. Let p : X Y
be defined by p([]) = (1).
ea homeomorphically onto U .
Lemma 10. p is continuous and open and maps U
Proof. *** Extra credit. ***

Lemma 11. Suppose a1 and a2 are classes of paths starting at y0 and having a
common endpoint y. Let U be a member of B such that y U . If a1 6= a2 , then
ea and U
ea are disjoint.
U
1
2
Proof. *** Extra credit. ***

Lemma 12. p is a covering map.


Proof. *** Extra credit. ***

Given a path starting at y0 and given t [0, 1], let t be the path given by
t (s) = (ts) for 0 s 1. You may find it helpful to describe by means of words
and/or pictures the path t 7 [t ] in X. [You dont need to write this down.]
Lemma 13. The path t 7 [t ] in X is indeed a path, i.e. it is continuous.
Proof. *** Extra credit. ***

Note that this lemma shows you how to lift any path in Y starting at y0 to a
path
in X starting at x0 := [ey0 ].
Lemma 14. X is path connected.

Proof. *** Extra credit. ***

Lemma 15. X is simply connected.


Proof. *** Your proof goes here. ***

[Hint: If [] 1 (Y, y0 ) is nontrivial, show that its lift


is not a closed curve;
deduce the result from this by covering space theory.]
Lemma 16. There is an action of 1 (Y, y0 ) on X that makes p a regular 1 (Y, y0 )cover.
Proof. *** Extra credit. ***

This completes the proof of Theorem 2.


Remark 2. It is possible to show that X covers every other path-connected covering
space of Y . [One can use Lemma 8.] For this reason X is often called the universal
cover of Y .
This is the end of the exam unless you want to read the next section
for fun.
7. Cayley graphs and the figure 8
This (optional) section assumes a little more knowledge of group theory than I
have assumed previously.
Let G be a group with a generating set S. The Cayley graph of (G, S) is the
graph = (G, S) with G as its vertex set and with an edge joining g and gs for
each g G and s S. For example, the Cayley graph of the free group F2 with
its standard generating set is the snowflake. A second example is shown in the
following picture; see if you can guess what G and S are. [Hint 1: The number of
vertices is the order of the group. Hint 2: Each generator s S is assigned a color;
so the number of colors of edges is the number of generators.]

The left-translation action of G on itself induces an action of G on the graph


and hence an action of G on the geometric realization X := ||. This action is
fixed-point free and makes X a regular G-cover of Y := X/G. If G = F2 with its
standard generating set, X is a tree and hence is contractible. In particular, it is
simply connected. And Y is the figure 8 in this case. Theorem 1 therefore implies
that the fundamental group of the figure 8 is free of rank 2, as claimed.

Mathematics 4530
Prelim 2
Brief solutions

1. Introduction
A crucial feature of the exponential covering p : R S 1 is that two points in the
same fiber differ by an integral multiple of 2. This was used in the definition of the
degree homomorphism 1 (S 1 ) Z, which was then proved to be an isomorphism.
We wish to generalize this to other covering maps, the so-called regular ones. As a
consequence, one can calculate many fundamental groups with no more effort than
was required for S 1 .
2. Group actions
We have defined a group to be a set G together with a binary operation
satisfying certain axioms. This may give you the wrong impression of group theory,
in that it gives you no clue as to how groups arise in nature. The way groups arise
in nature is that they act on things, thereby exhibiting the symmetry that those
things have.
Definition 1. Let G be a group and X a topological space. By an action of G
on X we mean a function G X X, denoted (g, x) 7 g x, satisfying:
(1) g (h x) = (g h) x for all g, h G and x X.
(2) 1 x = x for all x X, where 1 is the identity element of G.
(3) For each g G, the map x 7 g x is continuous.
It follows that the map x 7 g x, which we will call the action of g, is actually a
homeomorphism. Its inverse is given by the action of g 1 . Group actions are studied
extensively in algebra courses, such as Math 4340. For our purposes, however, it
suffices to have some examples.
Example 1. The additive group Z acts on R by (n, x) 7 2n + x for n Z and
x R. (We say Z acts on R by translation.) This is the action thats implicit in
our work on the exponential covering map, which led to the calculation 1 (S 1 )
= Z.
Example 2. The multiplicative group {1} acts on S 2 by scalar multiplication.
This action is related to the double covering S 2 P 2 mentioned in class, where
P 2 is the projective plane.
Example 3. More generally, whenever we have a 2-fold covering map p : X Y ,
there is an associated action of a group of order 2 on X. This is essentially what
you proved in additional problem 6 on Assignment 10.
Example 4. Let Zn be the additive group of integers mod n. There is an action of
Zn on S 1 , in which the integer k mod n acts as rotation through 2k/n radians.
Alternatively, if we replace Zn by the isomorphic group consisting of the nth roots
of unity in C, then we can describe the action using multiplication of complex
numbers.
1

Example 5. The concept of Cayley graph, which I will explain in Section 7, leads
to many examples of group actions. For example, the snowflake that covers the
figure 8 is the Cayley graph of the free group F2 ; as a result, there is an action of
F2 on the snowflake.
To simplify the notation in what follows, we will always use multiplicative notation for our groups (i.e., we will write gh instead of g h), and similarly we will
write (g, x) 7 gx for group actions.
Every one of the examples above has the property that it is fixed-point free, i.e.,
the action of any g 6= 1 has no points x such that gx = x. The significance of this
for us is that if two points x, x0 are in the same G-orbit, i.e., if there is a g G
such that gx = x0 , then the element g is unique. [If gx = x0 and hx = x0 , then
gx = hx = h1 gx = x = h1 g = 1 = g = h.] We can think of g as
something like the difference between x and x0 . For the case of Z acting on R as
in Example 1, it really is the difference (divided by 2).
3. Regular G-covers
Definition 2. Let G be a group. A regular G-cover is a covering map p : X Y
together with a fixed-point-free action of G on X whose orbits are the fibers of p.
Lets spell this out explicitly:
Lemma 1. Let p : X Y be a covering map with an action of a group G on X.
Then p is a regular G-cover if and only if it satisfies the following two conditions:
(a) p(gx) = p(x) for all g G, x X.
(b) If p(x) = p(x0 ) then there is a (unique) g G such that x0 = gx.
Sketch of proof. (a) says that every orbit is contained in a fiber. (b) [except for the
uniqueness part] says that, conversely, every fiber is contained in an orbit. So the
orbits are the fibers. It remains to sort out the uniqueness part of (b). We showed
above that if the action is fixed-point free, then uniqueness holds in (b). Conversely,
suppose the uniqueness assertion in (b) holds. If gx = x, then gx = 1x = g = 1,
so the action is fixed-point free.

Remark 1. The group action is useful in connection with path liftings. Namely, if
is a path in Y and
is the lift starting at some point in the fiber over (0), then
we can construct all the other liftings of by applying the group action. Indeed,
if
starts at x, let x0 be any other point in the same fiber. Then x0 = gx for some
g G, and g
is the lift of starting at x0 . Here (g
)(t) := g
(t) for t I; it is a
lift of by Lemma 1(a).
We have already seen many examples of regular G-covers. Indeed, every example
in the previous section has an associated regular G-cover, which was mentioned
explicitly except in Example 4. On the other hand, there are many covering maps
that are not regular because they do not have enough symmetry. Indeed, conditions
(a) and (b) imply, intuitively, that a regular G-cover looks the same no matter where
we stand in a given fiber. For example:
Lemma 2. Let p : X Y be a regular G-cover for some group G, and let be a
closed curve in Y that admits a closed lift
. Then every lift of is closed.
Sketch of proof. This is immediate from Remark 1, since g
is closed if
is closed.


One can deduce from this that not all covering maps are regular.
Example 6. In class we saw a 3-fold cover of the figure 8 by a union of 4 circles. It
was explicitly pointed out that there are closed curves in the figure 8 with a closed
lift at one point but a non-closed lift at another. So that cover is not regular.
4. The generalized degree homomorphism
Recall that we defined the degree of a closed curve in S 1 by lifting it to R and
measuring the difference between the starting and ending points of the lift. We can
do something similar for any regular cover.
Let p : X Y be a regular G-cover. Pick a basepoint y Y and a basepoint
x p1 (y). Given a path homotopy class [] 1 (Y, y), let
be the lift of
starting at x. Then
ends in the fiber p1 (y), so there is a unique g G such that

(1) = gx. We set


([]) := g.
In case p is the exponential covering of S 1 , this is precisely the degree defined in
class. Imitating the four steps that occurred in our calculation of 1 (S 1 ), we will
prove:
Theorem 1. Suppose that X is simply connected. Then
: 1 (Y, y) G
is an isomorphism.
The four steps are given in the following four lemmas, valid for any regular
G-covering.
Lemma 3. is well defined.
Sketch of proof. If two closed curves , at y are path-homotopic, then their lifts
starting at x end at the same point by the homotopy lifting theorem. It follows
that () = (), so is well-defined on path-homotopy classes.

Lemma 4. is a group homomorphism.
Sketch of proof. Given [], [] 1 (Y, y), let := . To compute the lift of
at x, we need to lift starting at x and then lift starting where the first lift

finished. Let
, be the lifts starting at x. Then
(1) = gx and (1)
= hx,
where g := () and h := (). The lift of starting at the end of
is g (see
This ends at g (1)

Remark 1), so the lift of starting at x is


g .
= ghx. Thus
() = gh = ()().

Lemma 5. If X is path connected, then is surjective.
Sketch of proof. Given g G, choose a path in X from x to gx. Composing with p,
we obtain a closed curve in Y whose lift
at x is the original path from x to gx.
Then ([]) = g.

Lemma 6. If X is simply connected, then is bijective.
Sketch of proof. Suppose ([]) = ([]). Then the lifts ,
starting at x end at
Composing
the same point. Since X is simply connected, it follows that
'p .
with p, we obtain 'p , i.e., [] = [].


5. Examples
Example 7. Combining Theorem 1 and Example 2, we deduce that the fundamental
group of P 2 is the group of order 2, since we know that S 2 is simply connected.
Example 8. Applying Theorem 1 to the regular Z Z covering of the torus by
R R, we deduce that the fundamental group of the torus is Z Z.
Example 9. We proved in class, using the Seifertvan Kampen theorem, that the
fundamental group of the figure 8 is the free group F2 on two generators. A different
proof can be obtained by applying Theorem 1 to the snowflake cover of the figure 8.
I will sketch how this can be done in Section 7.
6. Existence of regular G-covers
It is natural to ask how widely applicable Theorem 1 is. In this section we will
see that, in principle, Theorem 1 can be used to calculate the fundamental group
of every reasonable space. Here reasonable has a somewhat technical definition:
Definition 3. A space Y is said to be semilocally simply connected (SLSC) if
every point y Y has a neighborhood U such that the inclusion U , Y induces
the trivial homomorphism 1 (U, y) 1 (Y, y).
This condition may seem strange; but in fact, most spaces that one meets in
nature satisfy the even stronger condition of being locally contractible. Heres
more evidence that the SLSC condition is reasonable:
Proposition 1. If Y admits a simply connected covering space, then Y is SLSC.
Sketch of proof. Let p : X Y be a covering map with X simply connected. Given
y Y , let U be an evenly-covered neighborhood of y and let V be one of the slices
of p1 (U ). Then the inclusion i : U , Y factors as a composite
U V , X Y,
so i factors through the trivial group and is therefore trivial.

Our goal in this section is to prove the following theorem:


Theorem 2. Suppose Y is path connected, locally path connected, and SLSC. Then
there exists a group G and a regular G-cover p : X Y such that X is simply
connected.
[Necessarily, G will be isomorphic to 1 (Y ).]
The first step is to figure out what X has to look like.
Lemma 7. (a) If U satisfies the condition in Definition 3, then every smaller
neighborhood of y also satisfies the condition.
(b) Suppose Y is SLSC and locally path connected. Let B be the collection of pathconnected open sets U such that 1 (U ) 1 (Y ) is the trivial homomorphism
for some (or every) basepoint in U . Then B is a basis for Y .
(c) For any U B let 1 , 2 be paths in U with 1 (0) = 2 (0) and 1 (1) = 2 (1).
Then [1 ] = [2 ] in the fundamental groupoid of Y .

[Extra credit.]
Let Y be a path-connected space with basepoint y0 , let G be its fundamental
groupoid, and let G0 G be the subset consisting of path classes starting at y0 .
Thus G0 is a set whose elements are equivalence classes a = [], where is a path
in Y with (0) = y0 . Let p : X Y be a covering map, and let x0 p1 (y0 ) be a
basepoint. Then one can use path lifting to define a function h : G0 X. Namely,
given [] G0 , we lift to a path
in X starting at x0 and set
h([]) =
(1).
Lemma 8. (a) h is well defined.
(b) h is surjective if X is path connected and bijective if X is simply connected.
Sketch of proof. (a) Use the homotopy-lifting theorem. (b) Suppose X is pathconnected. Given x X, choose a path from x0 to x, and let be its image in Y .
Then
is the original path from x0 to x, hence h([]) = x and h is surjective.

Suppose X is simply connected. If h([]) = h([]), then


(1) = (1),
hence
'p .
This implies that 'p , i.e., [] = [].

From now on we assume that Y is path connected, locally path connected, and
SLSC. We have just seen that if Y admits a simply connected covering space X,
then the points of X must be in 11 correspondence with the elements of the set G0 .
With this as motivation, we will prove Theorem 2 by starting with G0 and putting
a suitable topology on it. Let B be the basis for Y given in Lemma 7(b).
ea G0 be
Let a = [] be an element of G0 . For any U B with (1) U , let U
the set of path classes of the form a b, where b = [] for some path in U with
ea a basic neighborhood of a, and we will call a subset
(0) = (1). We will call U
of G0 open if it contains a basic neighborhood of each of its points.
ea as
Lemma 9. The open sets just defined form a topology on G0 , with the sets U
a basis.
[Extra credit.]
As an aid to the intuition and a reminder of what were trying to do, we will
denote by X the set G0 with the topology that we have just defined. Let p : X Y
be defined by p([]) = (1).
ea homeomorphically onto U .
Lemma 10. p is continuous and open and maps U
[Extra credit.]
Lemma 11. Suppose a1 and a2 are classes of paths starting at y0 and having a
common endpoint y. Let U be a member of B such that y U . If a1 6= a2 , then
ea and U
ea are disjoint.
U
1
2
[Extra credit.]
Lemma 12. p is a covering map.

[Extra credit.]
Given a path starting at y0 and given t [0, 1], let t be the path given by
t (s) = (ts) for 0 s 1. You may find it helpful to describe by means of words
and/or pictures the path t 7 [t ] in X. [You dont need to write this down.]
Lemma 13. The path t 7 [t ] in X is indeed a path, i.e. it is continuous.
[Extra credit.]
Note that this lemma shows you how to lift any path in Y starting at y0 to a
path
in X starting at x0 := [ey0 ].
Lemma 14. X is path connected.
[Extra credit.]

Lemma 15. X is simply connected.


Sketch of proof. If [] 1 (Y, y0 ) is nontrivial, then the lift
of starting at x0
is given by t 7 [t ]. This ends at [], which is not x0 since [] is nontrivial. Thus

is not closed. This proves that the homomorphism p : 1 (X, x0 ) 1 (Y, y0 ) is


the trivial homomorphism. But we know that this homomorphism is injective, so
the domain 1 (X, x0 ) must be the trivial group.

Lemma 16. There is an action of 1 (Y, y0 ) on X that makes p a regular 1 (Y, y0 )cover.
[Extra credit.]
This completes the proof of Theorem 2.
Remark 2. It is possible to show that X covers every other path-connected covering
space of Y . [One can use Lemma 8.] For this reason X is often called the universal
cover of Y .

7. Cayley graphs and the figure 8


This (optional) section assumes a little more knowledge of group theory than I
have assumed previously.
Let G be a group with a generating set S. The Cayley graph of (G, S) is the
graph = (G, S) with G as its vertex set and with an edge joining g and gs for
each g G and s S. For example, the Cayley graph of the free group F2 with
its standard generating set is the snowflake. A second example is shown in the
following picture; see if you can guess what G and S are. [Hint 1: The number of
vertices is the order of the group. Hint 2: Each generator s S is assigned a color;
so the number of colors of edges is the number of generators.]

The left-translation action of G on itself induces an action of G on the graph


and hence an action of G on the geometric realization X := ||. This action is
fixed-point free and makes X a regular G-cover of Y := X/G. If G = F2 with its
standard generating set, X is a tree and hence is contractible. In particular, it is
simply connected. And Y is the figure 8 in this case. Theorem 1 therefore implies
that the fundamental group of the figure 8 is free of rank 2, as claimed.

Mathematics 4530

Final Exam

December 18, 2013

Do any 5 of the following 8 problems, and indicate clearly which 5 you have chosen.
You may use any result proved in class, in the assigned reading, or in the homework,
unless the problem asks you to prove such a result.
1. True or false. Justify each answer with a proof or counterexample.
(a) If X is a simply connected space and A is a path connected subspace, then A
is simply connected.
(b) If : I X is a path in a Hausdorff space X, then the image of is a closed
subset of X.
(c) Let X be a topological space and A a subspace, and let i : A , X be the
inclusion map. Then the induced homomorphism i : 1 (A, a0 ) 1 (X, a0 ) is
injective for any choice of basepoint a0 A.
(d) If U and V are path connected open subsets of a space X, then their intersection
is path connected.
e X be a covering map. Let x
e and let x0 :=
2. Let p : X
0 be a basepoint in X,
p(
x0 ). Suppose the following condition holds: For any closed curve in X at x0
such that [] is nontrivial in 1 (X, x0 ), its lift
starting at x
0 is not closed. Show
e x
that the group 1 (X,
0 ) is trivial.
3. Given a topological space X, we define the connected components of X to be
the equivalence classes determined by the following relation: x y if and only if
there is a connected subset of X containing both x and y.
(a) Prove that this relation is in fact an equivalence relation.
(b) Prove that every connected component is connected.
4. Let q : X Y be a quotient map. State and prove a universal mapping property
that characterizes continuous maps from Y to an arbitrary space Z.
5. Let X be a locally compact Hausdorff space whose one-point compactification is
metrizable. Show that X is the union of an increasing sequence of compact subsets
K1 K2 K3 .
6. (a) What does it mean to say that a topological space is regular ?
(b) What does it mean to say that a topological space is normal ?
(c) State Urysohns lemma and Tietzes extension theorem.
(d) Let X be a metric space and let (xn )n1 be an infinite sequence in X such that
d(xi , xj ) 1 for i 6= j. Prove that there is a continuous map f : X R such
that f (xn ) = n for all n.
7. (a) Define covering map.
e X is a covering map with X
e and X path connected. If X is
(b) Suppose p : X
simply connected, show that p is a homeomorphism.
8. (a) State the Seifertvan Kampen theorem, with hypotheses and conclusion
spelled out precisely.
(b) Give an example of a space whose fundamental group is a free group on 3
generators. Justify your answer as completely as you can; in particular, spell
out explicitly three closed curves that represent the 3 generators.

Very Brief Solutions to Final Exam


1. (a) False.
(b) True.
(c) False.
(d) False.
2. See the take-home prelim.
3. (a) Reflexivity and symmetry are obvious. Transitivity follows from the fact
that the union of two connected sets with nonempty intersection is connected.
(b) The connected component containing x is the union of all the connected sets
containing x.
4. See Munkres.
5. The point at infinity has a countable neighborhood base.
6. (a) See Munkres.
(b) See Munkres.
(c) See Munkres.
(d) The set A := {x1 , x2 , . . . } is closed and has the discrete topology. Apply the
Tietze extension theorem to the continuous map f : A R given by f (xn ) := n.
7. (a) See Munkres.
(b) p is an open surjection, so it suffices to prove that its injective. If p(
x1 ) = p(
x2 ),
choose a path
from x
1 to x
2 . Its image := p(
) is path homotopic to a
constant path, whose lifts are closed, so the lift
is also closed. Thus x
1 = x
2 .
8. (a) See Munkres.
(b) See additional problem 1 on assignment 13. Or consider a rose with three
petals.

You might also like