Notes
Notes
1. Introduction
What is geometry? Geometry is the study of rigid shapes that can be distinguished
with measurements (length, angle, area, . . . ).
What is topology? Topology is the study of shapes which are equivalent via deforma-
tions.
Topology versus Geometry: Objects that have the same topology do not necessarily
have the same geometry. For instance, a square and a triangle have different geometries
but the same topology.
Motivation: Our goal is to understand the shape of our universe. Consider the following
examples of two-dimensional universes: a plane, a sphere, a torus, and planes connected
by tubes.
These are topologically distinct universes. Intuitively, we can see that their “holes” distin-
guish them. Hence we seek a mathematical way to describe the holes; one familiar concept
we will use is continuity, which is related to a lack of holes.
3
4 1. DEFINITIONS AND EXAMPLES
Compare this to the usual definition of a metric space. The above definition is equivalent
to the usual definition of a metric space, but does not explicitly state the properties of
positivity or symmetry:
(1) d(a, b) ≥ 0 ∀a, b ∈ M (positivity)
(2) d(a, b) = d(b, a) ∀a, b ∈ M (symmetry).
Definition. The usual metric on Rn is
v
u n
uX
d((x1 , . . . , xn ), (y1 , . . . , yn )) = t (xi − yi )2 .
i=1
The discrete metric can be useful for testing conjectures as it does not rely on Rn .
Example (The Comb Metric for R2 .). Let X0 = {0} × [0, 1], Y0 = [0, 1] × {0}; and ∀n ∈ N,
let Xn = { n1 } × [0, 1]. Let M = (∪∞
n=0 Xn ) ∪ Y0 . The distance is the distance measured
along the comb in R2 .
...
0 1
3. CONTINUITY, METRIC SPACES, AND OPEN BALLS 5
We’d like to be able to talk about continuity in metric spaces (not just the reals).
Definition. Let (M1 , d1 ) and (M2 , d2 ) be metric spaces and a ∈ M1 . Let f : M1 → M2 .
We say f is continuous at a if ∀ > 0 there exists a δ > 0 such that ∀x ∈ M1 with
d1 (x, a) < δ then d2 (f (x), f (a)) < .
a
f(x)
δ
x ε
f(a)
Definition. Let (M,d) be a metric space and a ∈ M . The open ball of radius > 0
about a is defined to be
B (a) = {x ∈ M | d(x, a) < }
6 1. DEFINITIONS AND EXAMPLES
f(Bδ)
δ
ε
Bδ
It should be noted that open balls don’t always look like open balls...
Example. Let M = upper half of plane with usual distance
-1 1
Example (Comb metric). Under the comb metric, B 1 ((0, 1)) is just an interval along the
2
y-axis starting at (0, 1).
B2 ((0, 1) is everything on the comb below the line of slope −1 that connects points (0, 1)
and (1, 0). Since it’s an open ball, the line is not contained in the ball.
B½((0,1))
...
B2((0,1))
0 1
4. Open sets
Proof. Let a ∈ M and > 0. We want to show that B (a) is open. That is, we want
to show that ∀x ∈ B (a) there exists an r > 0 such that Br (x) ⊆ B (a).
Let x ∈ B (a) and let r < − d(x, a) and r > 0. We claim that Br (x) ⊆ B (a).
Let y ∈ Br (x). Then d(y, a) ≤ d(y, x) + d(x, a) < − d(x, a) + d(x, a) = .
Therefore, Br (x) ⊆ B (a)
8 1. DEFINITIONS AND EXAMPLES
Definition. Let
[
Ui = {x ∈ Ui | i ∈ I}
i∈I
Theorem 1 (the “open sets are nice” theorem). Let F be the family of open sets in a
metric space (M, d). Then:
(1) M , ∅ ∈ F
(2) If U, V ∈ F then U ∩ V ∈ F
S
(3) If ∀i ∈ I, Ui ∈ F then i∈I Ui ∈ F
Proof.
(2) Let x ∈ U ∪ V , let r1 > 0 such that Br1 (x) ⊆ U , and let r2 > 0 such that
Br2 (x) ⊆ V . Let r = min{r1 , r2 }. Then clearly Br (x) ⊆ Br1 (x) ⊆ U and
Br (x) ⊆ Br2 (x) ⊆ V . Thus Br (x) ⊆ U ∩ V as desired.
S
Let x ∈ i∈I Ui . We want to show that there exists an r > 0 such that Br (x) ⊆
(3) S
i∈I Ui .
S
Because x ∈ i∈I Ui , there exists an i0 ∈ I such that x ∈ Ui0 . Then there must
exist an r > 0 such that Br (x) ⊆ Ui0 . Then
[
Br (x) ⊆ Ui0 ⊆ Ui
i∈I
S
and hence i∈I Ui ∈ F .
Proof.
5. Topologies
Definition. Let X be a set with at least 2 points. Let F = {X, ∅}. Then we say (X, F )
is the indiscrete, or concrete topology.
Definition. If F1 and F2 are topologies on X and F1 ⊆ F2 then we say that F1 is weaker
than F2 , or F2 is stronger than F1 .
Two notes:
• weaker = fewer = coarser and stronger = more = finer.
• The discrete topology is the strongest topology on M and the indiscrete topology
is the weakest topology on X.
Example. Let X = R and U ∈ F iff U is the union of sets of the form [a, b) such that
a, b ∈ R. (This is called the half-open interval topology).
so F is stronger.
Example. Let X = R2 have the “dictionary order”. This means that (a, b) < (c, d) if
either a < c or a = c and b < d.
Note that a set can be both open and closed as well as neither open nor closed. (It is not
a door!)
S
Example. Let us consider R with the half-open interval topology. ThenS[0, ∞) = n∈N [0, n)
is open. On the other hand, the complement of this set is (−∞, 0) = n∈N [−n, 0), which
is also open. Thus this is a clopen set.
Example. A “vertical line” in R2 with the dictionary order is both open and closed. The
proof is left as exercise.
Lemma. Let (X, F ) be a topological space and A be the set of all closed sets in X. Then:
(1) X, φ ∈ A
T
(2) If C, D ∈ A, then C D∈A
T
(3) If Ci ∈ A for every i ∈ I, then i∈I Ci ∈ A
The proof is basically the same as that for open sets back in section 4.
Definition. Let (X, F ) be a topological space and A ⊆ X. Let {Uj | j ∈ J} be the set of
◦ S ◦
all open sets contained in A. Then we define A = Int(A) = j∈J Uj , and we say A is the
interior of A.
Small Fact. Let (X, F ) be a topological space and A ⊆ X. Then
◦
(1) A ⊆ A
◦
(2) A is open
◦
(3) If U ⊆ A is open, then U ⊆ A
◦
(4) A is open iff A = A.
◦ S ◦
Proof. (1) Since A = j∈J Uj and Uj ⊆ A for every j ∈ J, A ⊆ A.
◦ ◦
(2) By definition, A is a union of open sets, so A is open.
12 1. DEFINITIONS AND EXAMPLES
S ◦
(3) Since U is open in A, U ∈ {Uj | j ∈ J}. Therefore U ⊆ j∈J Uj = A.
◦
Note that this means that A is the “largest” open set in A.
◦ ◦
(4) Suppose that A = A. Then A is open by part (2), hence A is open.
◦
Conversely, suppose that A is open. Then A ∈ {Uj | j ∈ J}. Hence A ⊆ A. From
◦
(1), A = A.
Example. In the half-open interval topology on R, Int((0, 1]) = (0, 1). To prove this,
assume that 1 ∈ Int((0, 1]) and show that it leads to a contradiction.
Example. In the finite-complement topology on R, Int((0, 1]) = ∅. This follows from the
fact that R is infinite and all subsets of (0, 1] are finite.
Example. In the dictionary-order topology on R2 , Int([0, 1] × [0, 1]) = [0, 1] × (0, 1).
1 1
R Int(R)
1 1
Definition. Let A be a subset of a topological space (X, F ), and let {Fj | j ∈ J} be the set
T
of all closed sets containing A. Then the closure of A is defined as A = cl(A) = j∈J Fj .
Small Fact. Small facts about closures:
(1) A ⊆ A
(2) A is closed
(3) If A ⊆ C and C is closed, then A ⊆ C
(4) A = A if and only if A is closed.
Lemma (an Important Lemma). Let (X, F ) be a topological space and Y ⊆ X. Then p ∈ Y
if and only if for every open set U ⊆ X containing p, U ∩ Y 6= ∅.
This is nice, because sometimes it’s easier to work with closed sets than with open sets.
Definition. Let X and Y be topological spaces, and let f : X → Y .
14 1. DEFINITIONS AND EXAMPLES
This is sort of like continuity, except that we care about the image of sets instead of their
preimages.
Example time!
Example. Let F be the half-open topology on R, and define a function
f : (R, F ) → (R, usual) by f (x) = x
• Is f continuous? Yes! An open set in (R, usual) is a union of intervals of the form
(a, b). We know that f −1 (a, b) = (a, b), which is open in F .
• Is f open? No. Take any U = [a, b) ∈ F . Then f (U ) = [a, b), which is not open
in (R, usual).
• Is f closed? Also no, since f [a, b) = [a, b) is not closed in (R, usual).
7. Homeomorphisms
Define the centers of X and Y to be x and y, respectively. Fix some point a on the
boundary of X, and some b on the boundary of Y (that is, a ∈ ∂X and b ∈ ∂Y ).
Define a function f as follows:
• f (x) = y
• f (a) = b
• For t ∈ ∂X, look at the distance along the boundary from a to t. Then f (t) is a
point proportionally far along the boundary of Y .
• For s ∈ int(X), draw the ray connecting x and s. Let t be the point at which
this ray intersects ∂X. Now, in Y , draw the ray connecting y and f (s). Then s
is mapped to a point on this ray that is proportionally far from y.
7. HOMEOMORPHISMS 15
a
x y
b
X Y
t f(t)
s
a f(s)
x y
b
X Y
• Is this well-defined?
– Yes! There is always exactly one point in Y that can be mapped to be a
point in X 1.
• Is this a bijection?
– Yes! The inverse is defined identically, so it would make sense for this to be
a bijection. Also, consider the images of concentric squares centered on x
under f : they are mapped to disjoint concentric circles centered on y.
• Is f continuous and open?
1This works because both a square and a circle are convex shapes - for example, for all points p, q ∈ X,
the line pq that connects p and q lies entirely within X. This also implies that x and y didn’t actually have
to be the exact centers of X and Y respectively.
16 1. DEFINITIONS AND EXAMPLES
– Yes! Intuitively, it’s easy to see that an open set in X is mapped to an open
set in Y , and that the preimage of an open set in Y is open.
Question: Can we extend this to (some) non-convex regions?
Answer: Sure. Just divide up the non-convex region into smaller regions. Actually, a
subregion will work as long as there is some point in its interior such that any ray from
that point intersects the subregion’s boundary exactly once.
However, there are limits to this: for example, a donut is not homeomorphic to a circle.
This is hard to show; we’ll see a proof later.
There are also some homeomorphisms that we might find unsettling. For example, a
knot in R3 and the unit circle S 1 = {(x, y) ∈ R2 |x2 + y 2 = 1} are homeomorphic; the
argument works exactly like the one used for ∂X when showing that a square and circle
are homeomorphic (above). It turns out that, while the knot and circle are homeomorphic,
their complements are not.
Example. Define a function f : [0, 1) → S 1 by
f (t) = cos(2πt), sin(2πt)
This takes the interval and wraps it counterclockwise around the circle.
• Is this a bijection?
– Yes! The interval wraps once around the circle; one end is open, so there is
no overlap.
• Is f continuous?
– Yes! First we define an open ball in S 1 as B,S 1 (a) = {x ∈ S 1 | d(x, a) < }.
Then the preimage of every open ball in S 1 is an open interval, so f is
continuous.
• Is f open?
– No. Let U = [0, 1/2), which is open in [0, 1). Sadly, its image is not open in
S1.
So this f is not a homeomorphism.
8. Subspaces
9. Bases
You may have noticed that, in metric spaces, the idea of open sets was quite useful. We’d
like to extend this idea to more general topological spaces.
Definition. Let (X, F ) be a topological space and β ⊆ F such that for every U ∈ F , U
is a union of elements in β. Then we say that β is a basis for F .
Note that this basis is not necessarily minimal (like the basis of a vector space), but it is
more useful the more specific it is.
Example. R with the half-open interval topology was defined by the basis [a, b) | a < b .
Example. R2 with the dictionary topology was similarly defined by a basis of open inter-
vals.
Theorem 4. Let X be a set and β be a collection of subsets such that
S
(1) X = B∈β B
(2) ∀B1 , B2 , if x ∈ B1 ∩ B2 , then there is some B3 ∈ β such that x ∈ B3 ⊆ B1 ∩ B2 .
Let F = collection of elements of β. Then F is a topology on X with basis β.
1. Quotient Spaces
(2) (⇒) Suppose π is one-to-one. Let x, y ∈ X and x ∼ y. Then [x] = [y] and
π(x) = π(y), implying x = y since π is one-to-one.
(⇐) Suppose “∼” is “=”. Let x, y ∈ X such that π(x) = π(y). Then {x} = [x] =
[y] = {y}, and x = y.
π −1 (U ∩ V ) = π −1 (U ) ∩ π −1 (V ) ∈ FX
Since π −1 (Ui ) ∈ F∼ ∀i, the arbitrary union of such sets must also be open. Thus
by the above equality, π −1 (∪i∈I Ui ) ∈ F∼ , completing the proof.
Note that the continuity of π follows directly from the quotient topology.
Consider the interval [0,1] mapped to a circle under π. Then the interval [0,1/2), which is
open in [0,1] is mapped to a half circle which is not open in the whole circle.
π
(1,0)~(1,1)
(0,0)~(0,1)
(0,0) (1,0)
X X/~
Example. Let X = R2 and suppose (x, y) ∼ (x0 , y 0 ) if and only if ∃n, m ∈ Z such that
x = x0 + n, y = y 0 + m.
Note that X may be divided into integer side length squares, such that under ∼ all of the
given squares are equivalent. Thus we need only consider one such square, noting that the
opposite sides are equivalent, yielding a torus. This is pretty much the same as the above
example, except that we roll up the cylinder too. In the picture, we are identifying all lines
of the same color.
We may also generalize this idea to higher dimensions, yielding the analogous torus for
that dimension (i.e. X = R3 yields a 3-torus and so on).
24 2. MAKING NEW SPACES FROM OLD
Note that for n = 1 we obtain the unit circle such that points connected by a diameter are
equivalent. Thus any semi-circle forms a fundamental domain. Such a semicircle has it’s
endpoints as equivalent, and is thus topologically equivalent to the original circle.
We denote the quotient space S n / ∼ by RPn , or the real projective space. The above
argument concludes that RP1 ∼
= S1.
Example. From the above, RP2 = S 2 / ∼. We simply note that the fundamental domain is
a hemisphere whose boundary takes on the same topology as RP1 since it is a circle under
∼.
0 0
Example. Let X = R2 and define (x, y) ∼ (x0 , y 0 ) ⇔ x2 + y 2 = x 2 + y 2 .
Under ∼, any points on a circle centered at the origin are equivalent. Collapsing all such
circles to points, we find that X/ ∼ is a ray emanating from the origin (it’s pink in the
picture).
2
2
/~
Ff = {U ⊆ Y |f −1 (U ) ∈ FX }
Example. Let f map the closed segment [0, 1] to a figure-eight, where f (0) = f (1) = f ( 12 )
at the intersection of the two sides of the figure-eight. What is open in the quotient topology
(Y, Ff )?
• Any open-looking interval on the eight that does not contain the intersection point
is open since its preimage is clearly open on the segment; for the same reason, any
appropriate union or intersection of these open intervals is also open.
• Furthermore, any open set in the figure-eight that includes the point at the cross
must also contain an open interval of nonzero length extending along each of the
four legs of the figure-eight. This is necessary because the definition of Ff requires
that the preimage of anything open must be open itself; hence the only way for the
preimage of a set containing the intersection point to be open is if the preimage
is an open set in [0, 1] containing the three preimages of the intersection point (0,
1
2 , and 1).
-1
f
If a diagram commutes, then the path taken does not affect the result. Note that the arrow
representing g is dashed because we have not yet shown that the diagram commutes. In
the proof, we will define g so that the diagram does commute.
26 2. MAKING NEW SPACES FROM OLD
f
X5 /Y
55 D
55
55
55π g
55
55
5
X/ ∼
Proof. We want to show that there is some function g such that the following diagram
commutes:
f
A /B
πA πB
C _ _ _ g _ _ _/ D
For the past several pages, we’ve been building new topological spaces from old by using
equivalence relations to form quotient spaces. Here, we will change directions and build
new topological spaces from old by taking products of spaces. To that end, we wish to
define the product of two sets:
Definition. Let X, Y be sets. The product of X and Y is given by:
X × Y ≡ {(x, y) : x ∈ X, y ∈ Y }
This is a topology class, so our intrinsic urge is to find a natural topology for the product
of two topological spaces. While our first instinct might be to take products of open sets
in our original spaces, this approach will give unsatisfactory results:
Example. Consider the sets A = (0, 1) × (0, 1) and B = (1/2, 3/2) × (1/2, 3/2) as subsets
of R × R = R2 with the usual topology. Then A ∪ B in R2 is NOT a product of an open
set in R with an open set in R as we would like. To intuitively see that this is not the case,
see the picture! On the other hand, A ∪ B is open in the usual topology on R2 .
30 2. MAKING NEW SPACES FROM OLD
Although products of open sets will not work because they are not closed under unions,
we can use products of open sets to construct our topology. Define the following:
Definition. Let (X, Fx ) and (Y, Fy ) be topological spaces. Then the set βX×Y is defined
by:
βX×Y = {A × B : A ∈ Fx , B ∈ Fy }
and define:
FX×Y = {∪i∈I Ui : I is some index set, and Ui ∈ βX×Y }
In other words, βX×Y is the set of products of an open set in X and an open set in Y , and
FX×Y is the set of unions of elements in βX×Y . The motivation for defining our sets this
way is that we want FX×Y to be a topology on X × Y and for β to be its basis. We will
now verify this claim with the following small fact:
Theorem 9. If (X, Fx ) and (Y, Fy ) are topological spaces, the space (X × Y, FX×Y ) is a
topological space with basis βX×Y .
Proof. We will apply our basis theorem; i.e. we want to show that
S
(1) U ∈βX×Y U = X × Y
(2) Given B1 , B2 ∈ βX×Y , for each x ∈ B1 ∩ B2 , there exists B3 ∈ BX×Y such that
x ∈ B3 ⊆ B1 ∩ B2 .
For the first statement, we know that X ∈ Fx and Y ∈ Fy by definition of a topology so
that X × Y ∈ βX×Y . It follows then that because each U ∈ βX×Y is subset of X × Y :
[
X ×Y ⊆ U ⊆X ×Y
U ∈βX×Y
S
so that X × Y = U ∈βX×Y U , as desired.
For the second statement, let B1 , B2 ∈ βX×Y and let (x, y) ∈ B1 ∩ B2 . Then by definition
of βX×Y , there exist U1 , U2 ∈ Fx and V1 , V2 ∈ Fy such that B1 ∩ B2 = (U1 × V1 ) ∩ (U2 × V2 ).
Our aim is to find B3 ∈ βX×Y such that B3 contains (x, y) and B3 ⊆ B1 ∩ B2 , so define:
B3 = (U1 ∩ U2 ) × (V1 ∩ V2 ).
Thus U1 , U2 ∈ Fx ⇒ U1 ∩ U2 ∈ Fx by the closure of topologies under finite intersections
and similarly, V1 ∩ V2 ∈ Fy , so B3 ∈ βX×Y . Since (x, y) ∈ (U1 × V1 ) ∩ (U2 × V2 ), then
(x, y) ∈ (U1 × V1 ) and (x, y) ∈ (U2 × V2 ). Therefore, x ∈ U1 , U2 and y ∈ V1 , V2 , so
x ∈ U1 ∩ U2 , and y ∈ V1 ∩ V2 . It immediately follows by the definition of the intersection
of sets that:
(x, y) ∈ (U1 ∩ U2 ) × (V1 ∩ V2 ) = B3
The only thing left to show is that B3 ⊆ B1 ∩ B2 . To that end, let (a, b) ∈ B3 . Then
a ∈ U1 ∩ U2 and b ∈ V1 ∩ V2 by definition of B3 . It follows that:
a ∈ U1 , U2 and b ∈ V1 , V2 ⇒ (a, b) ∈ U1 ×V1 and (a, b) ∈ U2 ×V2 ⇒ a ∈ (U1 ×V1 )∩(U2 ×V2 ).
2. THE PRODUCT TOPOLOGY 31
We have successfully devised a topology for product spaces as unions of products of open
sets.
2.1. Examples.
Example. The set S 1 × [0, 1] looks like a cylinder! What kinds of sets are open in the
cylinder? A particular example is an open disc projected on the face of the cylinder and
unions thereof.
[0,1]
S1
Example. The set S 1 × S 1 looks like a torus! What kinds of sets are open in the torus?
Similar to the previous example, open discs projected onto the torus surface are examples
of open sets in S 1 × S 1 . In the picture, one copy of S 1 is red and one is green; they
determine a torus.
Not every space is a product of more than one space. For example the sphere S 2 is not
a product of more than one space. An intuitive (non-rigorous) justification is that the
natural axes of a sphere are the great circles; however, every pair of distinct great circles
intersect twice which makes it hard to define a coordinate system on the sphere.
2.2. The Product Projection Map. Now that we have a product topology to im-
pose on product spaces, we want a way to relate the product topology to the topologies of
the constituent spaces. In order to do so, we will define the projection maps as follows:
Definition. Let (X, Fx ) and (Y, Fy ) be topological spaces and create X × Y endowed
with the product topology FX×Y . Define πX : (X × Y, FX×Y ) → (X, Fx ) and πY : (X ×
Y, FX×Y ) → (Y, Fy ) by:
πX ((x, y)) = x πY ((x, y)) = y.
The map πX is the projection onto X and πY is the projection onto Y .
A small fact which we will derive now is that the product projection maps are continuous:
Theorem 10. Let (X, FX ) and (Y, FY ) be topological spaces and (X × Y, FX×Y ) be their
product with the induced product topology. Then the projection maps πX and πY onto X
and Y are continuous.
Recall that the quotient projection map is not necessarily an open map. It turns out that
the product projection map is an open map. Accidentally assuming that the quotient map
is open is a very common mistake that one should be aware of! We will now prove that
the product projection map is open:
Theorem 11. Let (X, FX ) and (Y, FY ) be topological spaces and (X × Y, FX×Y ) be their
product with the induced product topology. Then the projection maps πX and πY onto X
and Y are open.
2.3. Using Bases More Effectively. In metric spaces, open balls are bases for the
metric topology. By proving properties about open balls, we were able to say they apply
to the entire set. We would like to prove the following important yet tiny lemma so that
we can use bases in topological spaces like open balls in metric spaces:
Lemma. Let (X, Fx ) be a topological space with basis β. If W ⊆ X then W ∈ Fx if and
only if for all p ∈ W, there exists Bp ∈ β such that p ∈ Bp ⊆ W .
In other words, if we have a topological space with a basis, then every point in an open
set U is an element of a basis element contained in U , giving us a structure very similar to
the metric topology.
The following example illustrates how this lemma makes it very easy to define continuous
functions to the product space:
Example. Suppose f : R → R and g : R → R with f (x) = x2 + 3x and g(x) = sin(x).
Then the map h : R → R2 defined by h(x) = (x2 + 3x, sin(x)) is continuous because f, g
are.
This means that we can either look at A × B as a subspace of X × Y , then induce the
subspace topology, or we can induce the subspace topologies on A and B, then take the
cross product. Essentially, taking the cross product and creating subspaces “commute.”
Proof. We start by reviewing which sets are open under each topology:
FS = {(A × B) ∩ U | U ∈ FX×Y }
FX×Y = unions of sets of the form U × V , with U ∈ FA , V ∈ FB
2. THE PRODUCT TOPOLOGY 35
Let U ∩ (A × B) ∈ FS . Then
[
U ∩ (A × B) = (Ui × Vi ) ∩ (A × B) (where ∀i ∈ I, Ui ∈ FX , Vi ∈ FY )
i∈I
Then
[
U ∩ (A × B) = (Ui × Vi ) ∩ (A × B)
i∈I
= {(x, y) ∈ A × B | (x, y) ∈ Ui0 × Vi0 for some i0 ∈ I}
= {(x, y) ∈ X × Y | x ∈ Ui0 ∩ A, y ∈ Vi0 ∩ B for some i0 ∈ I}
[
= (Uj ∩ A) × (Vj ∩ B) for some index setJ}
j∈J
∈ FX×Y
The same argument run backwards shows that FX×Y ⊆ FS .
Therefore, FS = FX×Y .
2.6. Products and Quotient Spaces. Taking products and quotients does not
“commute” in the same way that taking products and subspaces does. We demonstrate a
(lengthy) counterexample to this idea.
Consider R with the usual topology, with x ∼ Y iff x = y or x, y ∈ N. This looks something
like a “sideways infinite flower.”
If a set open in R/ ∼ contains a (“the”) natural number, then its preimage must contain
every natural number, so it must contain an open interval about “the” natural number in
every direction (all infinity of them).
36 2. MAKING NEW SPACES FROM OLD
More formally, let π : R → R/ ∼ be the projection map. Let idQ : Q → Q be the identity
map, where each copy of Q has the usual topology. Note that idQ is a quotient map (where
∼ is ‘=’).
Now, consider π × idQ : R → Q → (R/ ∼) × Q by (π × idQ )(x, y) = (π(x), i(y)). We claim
that π × idQ is not a quotient map. To prove this, we show that the product topology on
(R/ ∼) × Q is not the same as the quotient topology
{U ∈ (R/ ∼) × Q | (π × idQ )(U ) ∈ FR×Q }
To do this, let’s find some U 0 ∈ Fπ×i such that U 0 ∈
/ F(R/∼)×Q . We want to construct U 0
as a union.
For all n ∈ N define Un to be the interior of the region of R × Q bounded
√ by the vertical
lines n − 41 and n + 14 and above and below by the lines through (n, n2 ) with slope ±1.
Each one of these sets has a vertical line through the natural number.
So ∀n ∈ N, {n} × Q ⊆ Un . Observe that ∀n ∈ N, Un ∈ FR×Q (it is an interior, so it is open).
S
Let n∈N Un ∈ FR×Q (also open in R × Q because it is a union).
Let U 0 = (π × i)(U ). This will glue together the strips along the vertical lines with R
coordinates in N. This can be pictured as a fan.
.
..
U1 U2 U3 U4 ...
U’
Claim: U 0 ∈ Fπ×i
Claim: U 0 6∈ F(R/∼)×Q .
2.7. Infinite Products. We would of course like to be able to generalize this to deal
with infinite products. Initially, we may simply want to define such products as follows:
X1 xX2 x...= {(x1 , x2 , ...)|xi ∈ Xi ∀i ∈ N}. However, such a definition limits us to countable
products, so we look more generally.
Definition. j ∈ J, let Xj be a set. Define the product Πj∈J Xj = {f : J → ∪j∈J Xj |f (j) ∈
Xj }. We refer to f (j) as the j th coordinate of the point f .
Example. Suppose J = {1, 2}. Then Πj∈{1,2} Xj = {f : {1, 2} → X1 ∪ X2 |f (j) ∈ Xj } =
{(f (1), f (2))|f (j) ∈ Xj } = {(x1 , x2 )|xj ∈ Xj } = X1 x X2 . Thus we see that this definition
agrees with our previous definition for finite products.
38 2. MAKING NEW SPACES FROM OLD
Example. Consider Πj∈R {1, 2}. By definition this is equivalent to {f : R → {1, 2}|f (j) ∈
{1, 2}, j ∈ R}, which precisely correspond to subsets of R if we simply think of the preimage
of 1 under f as the elements in the set and the preimage of 2 under f as those elements
outside the set. Commonly this product is then denoted by {1, 2}R .
Now we wish to define a topology on these products which, as in example 1, agrees with
our prior definition for a product of two sets if the indexing set J = {1, 2}. Perhaps the
most natural way of doing this is to define the following basis:
This topology is not what we desire, but is a topology, aptly named the box topology on
the product.
Example. The product topology on Πj∈J Xj is given by the basis βΠ = {Πj∈J Uj |Uj = Xj
for all but finitely many j and ∀j ∈ J, Uj ∈ Fj }.
Remarks:
(1) βΠ ⊆ β
(2) Both are bases for topologies on the product.
Definition. Define the projection map πj : Πj∈J Xj → Xj by π(f ) = f (j).
Lemma. For all j ∈ J, Suppose (Xj , Fj ) is a topological space. Then πj as defined above
is continuous for all j ∈ J.
(⇐) Suppose that gj is continuous for all j ∈ J. Let U ∈ βΠ . We wish to show that
h−1 (U ) ∈ FY .
Note that h−1 (U ) = {y ∈ Y |h(x) ∈ U } As an open set in the product topology,
U = Πj∈J Uj where Uj ∈ Fj . Thus, h−1 (U ) = {y ∈ Y |h(x) ∈ Πj∈J Uj } = {f ∈
Πj∈J Uj such that ∀j ∈ J, f (j) = gj (y)} = {y ∈ Y |gj (y) ∈ Uj } = {y ∈ Y |y ∈
gj−1 (Uj )∀j ∈ J} = ∩j∈J Uj .
Since gj is continuous for all j, gj−1 (Uj ) is open in Y for all j. Also, by definition
of the product topology, Uj = Xj for all but at most finitely many j. Thus,
gj−1 (Uj ) = Y for all but at most finitely many j. It follows that the intersection
∩j∈J Uj is a finite intersection of open sets since removing all trivial indices will
not change the intersection. Thus, h−1 (U ) ∈ FY so h is continuous, completing
the proof.
CHAPTER 3
Distinguishing Spaces
Proof. Let y ∈ Y . Since f is surjective, there exists x ∈ X such that f (x) = y. Since
{x} ∈ FX and f open, f ({x}) ∈ FY . Thus all singletons in are elements of FY and all sets
in Y are unions of singletons and hence elements of FY , so every set is an open set and FY
is the discrete topology.
Lemma (Daniel’s Lemma). Let (X, FX ) and (Y, FY ) be topological spaces and f : X → Y be
a continuous bijection. If FX is the indiscrete topology, then FY is the indiscrete topology.
Unfortunately, even with these five lovely Topological Properties, we can’t yet distinguish
a circle from a line. Clearly there is more work to do.
Example (a non-example). Distance is not a topological property. A big circle is homeo-
morphic to a little circle.
1. Compactness
IMPORTANT WARNING!
We learned in Math 131 that in Rn , a set if compact ⇐⇒ it is closed and bounded. THIS
IS NOT TRUE IN TOPOLOGICAL SPACES. DO NOT TRY TO USE IT.
Example. Is the set (0, 1) compact in
(1) R with the discrete topology?
(2) R with the half-open topology?
(3) R with the finite complement topology?
Answers:
(1) No. Take the open covering {B1/2 (x) : x ∈ (0, 1)}. If we remove any of these
balls, the corresponding x will no longer be covered. However, there are clearly
uncountably infinitely many balls.
(3) Yes. Suppose we have a cover of (0, 1). Take any element of the cover, say
U47 . U47 is missing at most finitely may elements of (0, 1) since its complement
is finite. For each element x ∈ R \ U47 , select one Ujx containing x. The set
{U47 } ∪ {Ujx : x ∈ R \ U47 } is a finite open subcover.
Theorem 12 (analogous to a theorem from 131). Let (X, FX ) and (Y, FY ) be topological
spaces, S ⊆ X, and f : X → Y continuous. If S is compact then f (S) is compact.
Proof. Suppose that for some such S ⊆ X, no such p ∈ X exists. Therefore, [ for all
p ∈ X there exists some Up ∈ FX such that p ∈ Up and |S ∩Up | is finite. So, as Up = X,
p∈X
{Up | p ∈ X} is an open cover of X. As X is compact, there exists some finite subcover
n
[
{Up1 , ..., Upn }. Hence, from the definition of a subcover, X = Upi . Intersecting both
i=1
sides with S gives that:
n n
!
[ [
S= Upi ∩S = (Upi ∩ S)
i=1 i=1
However, as |Up ∩ S| is finite for all pS∈ X, |Upi ∩ S| is finite for 1 ≤ i ≤ n. The union
of finite sets is finite, and therefore | ni=1 (Upi ∩ S)| is finite. However, S is infinite by
assumption. So this is a contradiction, and no such S exists.
Theorem 16 (Finite Tychonoff Theorem). Let (X, FX ) and (Y, FY ) be compact topological
spaces. Then (X × Y, FX×Y ) is also a compact topological space.
(a) Consider some open cover. (b) Restrict to some {x} × Y , which is compact,
so has a finite subcover.
1. COMPACTNESS 45
Ux Ux Ux
1 2 3
(c) So there is some open Ux with x ∈ Ux s.t. (d) So consider just the Ux , which are a cover of
Ux × Y is contained in this subcover. X, so have a finite subcover.
So the original sets that covered {x} × Y corresponding to the chosen Ux for the desired
finite cover of X × Y .
2. Hausdorffness
This is continuous because the domain has the discrete topology, but the domain is Haus-
dorff while the range clearly is not.
Small Fact. Suppose f : (X, FX ) → (Y, FY ) is a homeomorphism and (X, FX ) is Haus-
dorff. Then (Y, FY ) is also Hausdorff.
2. HAUSDORFFNESS 47
Proof. Let p 6= q ∈ Y . Because f is a bijection, f −1 (p) and f −1 (q) are distinct points
in X.
Since X is Hausdorff, ∃ U, V ∈ FX such that U ∩ V = ∅, f −1 (p) ∈ U , f −1 (q) ∈ V .
Consider f (U ), f (V ). Since f is a homeomorphism and, thus, open, f (U ) and f (V ) are
open. Then, because f is a bijection, we have p ∈ f (U ), q ∈ f (V ), and f (U ) ∩ f (V ) =
∅
Flapan Says... “ Hausdorff-ness and compactness are like two people who love each other,
because when two people love each other, they can do so much more together than they
can alone.”
Proof. (Compare the following to the proof that all compact subsets of metric spaces
are closed. That is, we are going to show that we don’t need a metric, just Hausdorff-ness,
for closed subsets to be compact.)
Rather than show A is closed, we will show X − A is open.
Let p ∈ X − A. We want to show ∃ U ∈ FX such that p ∈ U ⊆ X \ A. Let a ∈ A. Because
X is Hausdorff, ∃ Ua , Va ∈ FX such that p ∈ Ua , a ∈ Va , Ua ∩ Va = ∅.
Consider {Va |a ∈ A}. This is an open cover of A because Va open and a ∈ Va ∀a ∈ A. So,
because A is compact, ∃ finite subcover {Va1 , Va2 , ..., Van }
Let U = ni=1 Uai . U ∈ FX since it is a finite intersection of elements of FX and p ∈ U
T
since p ∈ Uai ∀ i = 1, 2, ..., n.
Claim: U ⊆ X − A
Proof. ∀ i = 1, 2, ... n, we have Uai ∩ Vai = ∅. Also, A ⊆ ni=1 Vai and U ∩ A ⊆ U ∩ ni=1 Vai .
S S
However, U ∩ ni=1 Vai = ∅ because if ∃ x ∈ U ∩ ni=1 Vai , then ∃ io such that x ∈ Vaio ∩Uaio .
S S
But, by definition, Vaio ∩ Uaio = ∅, so no such x exists. Thus,
n
[
U ∩A⊆U ∩ Vai = ∅,
i=1
meaning U ∩ A = ∅. So, because U ⊆ X but U ∩ A = ∅, U ⊆ X − A
Thus, we have found U ∈ FX such that p ∈ U , U ⊆ X − A. Thus, X − A is open. Thus,
A is closed.
Corollary. In any Hausdorff space, (finite sets of ) points are closed sets.
48 3. DISTINGUISHING SPACES
Proof. Let p ∈ (X, FX ) and let (X, FX ) be Hausdorff. Let {Ui |i ∈ I} be an open
cover of {p}. Then ∃ io ∈ I such that p ∈ Uio . Thus {Uio } is a finite subcover. Hence, {p}
is compact.
Thus, by Theorem above, {p} is closed.
Lemma (it’s important!). Let f : (X, FX ) → (Y, FY ) be continuous. Let X be compact,
and Y be Hausdorff. Then f is a homeomorphism if and only if it is a bijection.
X X/~
Points of the same color are identified. Points on the x-axis are not separable, so this is a
Hausdorff space X and an equivalence relation ∼ such that X/ ∼ is not Hausdorff.
Example. Let X = R, and set x ∼ y if and only if x = y or x, y > 0.
Let p > 0. Then [p], [0] ∈ X/ ∼. Let U ⊆ X/ ∼ be an open set containing [0]. We claim
that [p] ∈ U . To see why, note that π −1 (U ) is an open subset of R containing 0. Hence,
3. NORMALNESS 49
Note: this proves that the continuous images of a Hausdorff space need not be Hausdorff!
3. Normalness
Definition. Let (X, Fx ) be a topological space. We say X is normal if for every pair of
disjoint closed sets A, B ⊆ X, there exist disjoint open sets U, V ⊆ X such that A ⊆ U
and B ⊆ V .
In particular, we have shown that a space can be Hausdorff but not normal. Under what
conditions must a Hausdorff space be normal?
Lemma. If (X, Fx ) is Hausdorff and compact, then X is normal.
A
a a
(a) Let A and B be closed and disjoint sets. (b) Each point in B can be separated from a by an
Let a ∈ A. open set. This forms an open cover. Take a finite
subcover.
(c) Repeat for each a ∈ A. This makes an open (d) Take the intersection of all of the covers of B.
cover of A and a lot of finite open covers of B. It’s disjoint from the cover of A, so A and B are
separated by open sets, so X is normal.
3. NORMALNESS 51
Proof. Let A and B be disjoint closed subsets of X. Then A and B are compact
because X is compact. Let a ∈ A. For every b ∈ B, there exist open sets Ub and Vb such
that a ∈ Ub and b ∈ Vb , and Ub ∩ Vb = ∅. Then {Vb | b ∈ B} is an open cover of B. Since
B is compact, we can choose a finite subcover {Vb1 , . . . , Vbn }. Let Ua = ∩ni=1 Ubi . For every
a ∈ A, Ua ∈ Fx and a ∈ Ua . Let Va = ∪ni=1 Vbi . Then B ⊆ Va ∈ Fx .
We claim that Ua ∩ Va = ∅ for all a ∈ A. To see why, note that (∩ni=1 Ubi ) ∩ (∪ni=1 Vbi ) = ∅
because for all i, Ubi ∩ Vbi = ∅.
Now, {Ua | a ∈ A} is an open cover of A. Since A is compact, this cover has a finite
subcover {Ua1 , . . . , Uam }. Now, V = ∩m i=1 Vai is open and B ⊆ V . For all i = 1, . . . , m,
Uai ∩ (∩m V
i=1 ai ) = ∅ because for all i = 1, . . . , m, Uai ∩ Vai = ∅.
Let U = ∪m
i=1 Uai ∈ Fx . Then A ⊆ U and U ∩ V = ∅ because Uai ∩ Vai = ∅ for all i.
Therefore, X is normal.
Theorem 18. Let (X, FX ) be a compact and Hausdorff topological space. If (Y, FY ) is a
topological space and f : X → Y is continuous, onto, and closed, then (Y, FY ) is compact
and Hausdorff.
4. Connectedness
• (U ∩ yj0 ) ∩ (V ∩ yj0 ) = ∅,
• (U ∩ yj0 ) and (V ∩ yj0 ) are open in yj0 ,
• yj0 is connected, so either U ∩ yj0 = ∅ or V ∩ yj0 = ∅.
Without loss of generality, say U ∩ yj0 = ∅. Then V ∩ yj0 = yj0 . By the same argument,
∀j ∈ J, yj ⊆ U or yj ⊆ V .
\ \
But yj 6= ∅, so there exists p ∈ yj such that p ∈ yj0 ⊆ V . Thus ∀j ∈ J, yj ⊆ V .
j∈J j∈J
[
Since X = yj ⊆ V , U, V is not a separation of X. Thus, X is connected.
j∈J
Theorem 20. Let (X, FX ) and (Y, FY ) be topological spaces. Then (X × Y, FX×Y ) is
connected if and only if both X and Y are connected.
The axes are connected, because X and Y are connected. If we make a copy of
the axes, shifted over a little bit, that copy is connected too. The union of these
shifted axis crosses will be the entirety of X × Y , and their intersection will be
nonempty, so X × Y is connected.
Now for the actual proof. Let (x0 , y0 ) ∈ X ×Y. Since {x0 }×Y ∼
= Y , it is connected
(as connectedness is a top. prop.). Similarly, {y0 } × X is connected.
Observe that (x0 , y0 ) ∈ ({x0 }×Y )∩({y0 }×X). That is, ({x0 }×Y )∩({y0 }×X) 6= ∅.
Then (x0 , y0 ) ∈ ({x0 } × Y ) ∪ ({y0 } × X) is connected by the Important Flower
Lemma. By the same process, we see that ∀x ∈ X, (X × {y0 }) ∪ ({x} × Y ) is
connected.
\
Note that (X × {y0 }) ⊆ (X × {y0 }) ∪ ({x} × Y ).
x∈X
4. CONNECTEDNESS 55
[
Claim: (X × {y0 }) ∪ ({x} × Y ) = X × Y .
x∈X
Proof of claim:
[
(⊆) Clearly (X × {y0 }) ∪ ({x} × Y ) ⊆ X × Y because X × Y is the entire
x∈X
space.
(⊇) Let (x1 , y1 ) ∈ X × Y . [
Then (x1 , y1 ) ∈ ({x1 } × Y ). Thus, ∀x ∈ X and y ∈ Y ,
(x, y) ∈ ({x} × Y ) ⊆ (X × {y0 }) ∪ ({x} × Y ).
x∈X
\
Since (X×{y0 })∪({x}×Y ) is connected ∀x ∈ X, and (X × {y0 }) ∪ ({x} × Y ) 6=
x∈X
∅, X × Y is connected by the Important Flower Lemma.
X X/~
Connectedness is not always intuitive, though. Here’s a large example showing this.
Define the Comb, the Flea, and the space X as follows.
∞
[
Comb: yn in R2 where y0 = [0, 1] × {0} and ∀n ∈ N, yn = {1/n} × [0, 1].
n=0
5. Path-Connectedness
Some remarks:
(1) Connected is a negative definition and path-connected is a positive definition (for
connectedness, we are trying to show that a separation doesn’t exist, so it is
easiest to do connectedness proofs by contradiction. For path-connectedness, we
are trying to show that a path exists, so path-connectedness proofs are more easily
done constructively).
(2) In general, it is easier to prove that a space is disconnected than to prove that it
is not path-connected.
(3) In general, it is easier to prove that a space is path-connected than to prove that
it is connected.
Theorem 21. If (X, FX ) is path-connected, then (X, FX ) is connected.
contradiction, since we know from Math 131 that [0, 1] is connected. Thus, (X, FX ) must
be connected.
Recall the example of the flea and the comb. We showed that the flea and comb is con-
nected.
Theorem 22. The flea and comb is not path-connected.
Proof. We want to show that @ a path from the flea to (0, 0).
Suppose ∃ a path f from the flea to (0, 0). Let p = flea. Then f −1 ({p}) is not empty
because f (0) = p by the definition of f . Similarly, by the definition of f , f (1) = (0, 0) 6= p,
so f −1 ({p}) is proper as well. Observe that since {p} is closed in X and f is continuous,
f −1 ({p}) is closed.
We’d like to show that f −1 ({p}) is open, since then it would be a nontrivial clopen subset
of the flea and the comb. This implies that the set is not connected, a contradiction.
Let y ∈ f −1 ({p}) be given.
Observe that B 1 (p; X) is open in X. Since f is a path, f is continuous, so f −1 (B 1 (p; X))
2 2
is open in [0, 1]. Since y ∈ f −1 ({p}), f (y) = p ∈ B 1 (p; X) ⇒ y ∈ f −1 (B 1 (p; X)). Thus,
2 2
since f −1 (B 1 (p; X)) is open in [0, 1], ∃ > 0 s.t. B (y; [0, 1]) ⊆ f −1 (B 1 (p; X)).
2 2
Let z ∈ B (y; [0, 1]). If f (z) = p, then B (y; [0, 1]) ⊆ f −1 (p) and so f −1 (p) is open,
completing the proof.
Suppose f (z) 6= p. Since z ∈ B (y; [0, 1]), z ∈ f −1 (B 1 (p; X)), so f (z) ∈ B 1 (p; X) (so
2 2
d(f (z), p) < 12 ). Since Y0 = [0, 1] × {0}, for each q ∈ Y0 , d(p, q) ≥ 1 ⇒ q 6∈ B 1 (p; X). Thus,
2
f (z) 6∈ Y0 . Thus since f (z) 6= p, ∃n ∈ N s.t. f (z) ∈ Yn .
There exists r ∈ R \ Q such that 0 < r < n1 . Define sets A, B ⊆ f (B (y; [0, 1])), as
A = {(x, y) ∈ f (B (y; [0, 1])) |x < r} and B = {(x, y) ∈ f (B (y; [0, 1])) |x > r}
We next show that A and B are a separation for f (B (y; [0, 1])). To this end, we’ll need
to show that A and B are disjoint, proper and non-empty, that their union is the entire
set, and that they’re are both clopen.
We note that A ∩ B = ∅ by definition. We know that f (z) ∈ B, and p ∈ A, so neither set
will be empty and both will be proper.
Next, we want to show that A ∪ B = f (B (y; [0, 1])). Certainly A ∪ B ⊆ f (B (y; [0, 1])).
1
Now let (x, f (x)) ∈ f (B (y; [0, 1])). We know that f (x) 6= 0, and in fact x = m for some
m ∈ N. Then x 6= r, so x ∈ A ∪ B and we’re done.
Next we want to show that A is open in f (B (y; [0, 1])). We know that {(x, y)|x < r} is
open in R2 , the half plane to the left of r. It follows using the subspace topology that
5. PATH-CONNECTEDNESS 59
A = {(x, y)|x < r} ∩ f (B (y; [0, 1])) is open in f (B (y; [0, 1])). Similarly, B is open in
f (B (y; [0, 1])). Since each is the other’s complement in f (B (y; [0, 1])), then both are also
closed.
We’ve hence shown that A and B are a separation of f (B (y; [0, 1])). This is a contradiction,
since B (y; [0, 1]) is connected and f is continuous, and thus we’ve disconnected B (y; [0, 1]).
We conclude that f (z) = p, for all such z ∈ B (y; [0, 1]).
Therefore B (y; [0, 1]) ⊆ f −1 ({p}), so f −1 ({p}) is open. So f −1 ({p}) is a clopen, non-
empty, proper subset of [0, 1]. This is a contradiction, so we conclude that there does not
exist a path in X from p to (0, 0) (or anywhere).
The point of the whole example is to show that path-connected is stronger than connected.
We knew already that path-connected implies connected, and now we see that connected
doesn’t necessarily imply path-connected.
5.1. Combining paths. It turns out that it’s useful to have a way to combine paths.
Definition. Let f, g be paths in a topological space (X, FX ) such that f (1) = g(0). Then
we define f ∗ g : I → X by
(
f (2t) 0 ≤ t ≤ 21
(f ∗ g)(t) =
g(2t − 1) 12 ≤ t ≤ 1
Intuitively what’s happening is that we’re connecting two paths while speeding things
up, creating a new single path parametrized from 0 to 1 from two paths which were
parametrized from 0 to 1.
Small Fact. f ∗ g is a path from f (0) to g(1).
Proof. By the Pasting Lemma1, since [0, 12 ] and [ 12 , 1] are closed subsets under [0, 1],
and f (2( 12 )) = f (1) = g(0) = g(2( 12 ) − 1), f ∗ g is continuous. So f ∗ g is a path. Since
(f ∗ g)(0) = f (0) and (f ∗ g)(1) = g(1), then f ∗ g is a path from f (0) to f (1).
Proof. The proof is identical to the connected one; the key is the flower lemma.
Definition. Let (X, FX ) be a topological space, and p ∈ X. Let {Cj |j ∈ J} be the
set of all path-connected subspaces of X containing p. Then ∪j∈J Cj is said to be the
path-connected component Cp .
6. Homotopies
We now have enough background to start talking about geometric (or algebraic) topology.
A few remarks:
(1) f is a path, because it is a composition of continuous functions.
(2) f is a path from f (1) to f (0).
(3) f ∗ f 6= the “identity”. We haven’t created an inverse.
Definition. Let (X, FX ) be a topological space, and a ∈ X. We define ea : I → X by
ea (t) = a for all t ∈ I.
This is probably the most important concept in the rest of these notes.
6. HOMOTOPIES 61
f0 f1
It is visually apparent that one image can be warped to make the other, so f0 and f1 are
homotopic. However, homotopic functions need not be homeomorphic.
f0 f2
Observe that F is continuous. Note also that F (x, 0) = (x, 0) = f0 (x) and F (x, 1) =
(x, x2 ) = f1 (x).
Thus F is a homotopy.
62 3. DISTINGUISHING SPACES
f1(1)
f1(0) f1(s) f1(1)
t=1
f1(s)
f1(0)
F(s0,t)
f0(1)
F(s0,t)
s0
f0(s)
t=0 s0 f0(1)
f0(0) f0(s) f0(0)
The image of a vertical segment in the square is the path taken by the x coordinate of that
segment during the homotopy.
A few remarks.
(1) Suppose Y is not path connected and f0 (x) and f1 (x) are in different path com-
ponents. Then there does not exist a homotopy from f0 to f1 .
f1 f1
F(0,t)
f0 f0
If f0 and f1 were homotopic, then F (0, t) would be a path from f0 (0) to f1 (0).
But f0 and f1 are in different path components.
(2) If Y is path connected and f0 and f1 are paths in Y , then f0 ' f1 . Homotope (the
verb!) f0 to its initial point, move it to the initial point of f1 and then stretch it
6. HOMOTOPIES 63
f0
f1 f1 f1 f1
F(s,0) b
a F(s,1)
Claim: f0 ∼ f1 .
Proof. Let F (s, t) = (1 − t)f0 (s) + tf1 (s) (Note: we call this the straight line
homotopy). Observe that ∀s ∈ I, F (s, 0) = f0 (s) and F (s, 1) = f1 (s), and that F is
continuous, so F is a homotopy from f0 to f1 . Now let t ∈ I be given. Observe that
F (0, t) = (1 − t)f0 (0) + tf1 (0) = f0 (0) = a, and that F (1, t) = f0 (1) = b. Thus ∀t ∈ I,
F (0, t) = a and F (1, t) = b, so F is a path homotopy, and thus f0 ∼ f1 .
Example. Let X ∼ 2
= D , let a, b ∈ X, and let f0 and f1 be paths in X from a to b. Then
f0 ∼ f1 .
Note that the same proof (using only the parts related to continuity and homotopy) works
for '.
Definition. Let (X, FX ) be a topological space and let a, b ∈ X be given. For each path
f from a to b in X, define [f ] to be the path homotopy class of f.
Some remarks:
(1) [f ], [g], and [f ∗ g] are not elements of the same quotient ‘world’ unless a = b = c.
Lemma (Important). The product of path homotopy classes is well-defined. Formally, let
(X, Fx ) be a topological space. Let f, f 0 be paths in X from a to b and let g, g 0 be paths in
X from b to c. Then:
f ∗ g ∼ f 0 ∗ g 0 ⇒ [f ][g] = [f 0 ][g 0 ]
a f’ b g’ c
F G
a b g c
f
Define H : I × I → X by:
(
F (2s, t) s ∈ [0, 1/2]
H(s, t) =
G((2s − 1), t) s ∈ [1/2, 1]
We claim H is a homotopy from f ∗ g to f 0 ∗ g 0 . We will now verify the claim.
We see that:
F (2 · (1/2), t) = F (1, t) = b G(2 · (1/2) − 1, t) = G(0, t) = b
so by the pasting lemma, it follows that H is continuous.
We now show H gives the desired paths by calculating:
H(0, t) = F (0, t) = a H(1, t) = G(1, t) = c
so H(s, [0, 1]) gives a set of paths from a to c.
Finally, we need to prove that H gives a homotopy between the intended paths:
( (
F (2s, 0) s ∈ [0, 1/2] f (2s) s ∈ [0, 1/2]
H(s, 0) = = =f ∗g
G(2s − 1, 0) s ∈ [1/2, 1] g(2s − 1) s ∈ [1/2, 1]
Similarly:
( (
F (2s, 1) s ∈ [0, 1/2] f 0 (2s) s ∈ [0, 1/2]
H(s, 1) = = 0
= f 0 ∗ g0
G(2s − 1, 1) s ∈ [1/2, 1] g (2s − 1) s ∈ [1/2, 1]
as desired.
We have shown that the product of two path homotopy classes is well defined. For the
purposes of defining a group of path homotopy classes, we would like a set of paths whose
products have the same endpoints as the original paths. This simplification motivates the
two definitions which follow:
Definition. Let f be a path in X such that f (0) = f (1) = x0 ∈ X. Then f is a loop in
X based at x0 .
Note that if f, g are loops in X based at some point x0 ∈ X, then their product f ∗ g is
also a loop based at x0 . In particular, we then have that [f ], [g] and [f ][g] = [f ∗ g] are all
path-homotopic classes of loops based at x0 .
Definition. Let (X, Fx ) be a topological space, and let x0 ∈ X.
Define π1 (X, x0 ) as the set of path homotopy classes of loops based at x0 endowed with
the product of path homotopy classes. We call π1 (X, x0 ) the fundamental group of X
based at x0 .
7.1. The Fundamental Group is a Group. Our ultimate goal is to harness the
power of group theory from abstract algebra to study topological spaces. We defined the
fundamental group of X based at x0 with the path homotopy class product operation.
Naturally, we would like to prove that π1 (X, x0 ) endowed with the product operation is
actually a group. In other words, if (X, Fx ) is a topological space with x0 ∈ X, we must
prove the following:
(1) π1 (X, x0 ) is closed under the operation [][]. In other words, for [f ], [g] ∈ π1 (X, x0 ),
[f ][g] ∈ π1 (X, x0 ).
(2) The operation [][] is associative. In other words, given [f ], [g], [h] ∈ π1 (X, x0 ):
([f ][g])[h] = [f ]([g][h])
(3) π1 (X, x0 ) contains an identity element. In other words, there exists [e] ∈ π1 (X, x0 )
such that for all [f ] ∈ π1 (X, x0 ):
[e][f ] = [f ][e] = [f ]
(4) Every element of π1 (X, x0 ) has an inverse. In other words, given f ∈ π1 (X, x0 ),
there exists g ∈ π1 (X, x0 ) such that:
[f ][g] = [g][f ] = [e]
Lemma (Closure). Let (X, Fx ) be a topological space, and let x0 ∈ X. Let [f ], [g] ∈
π1 (X, x0 ). Then:
[f ][g] ∈ π1 (X, x0 )
Next, we will show that path homotopy path products of loops based at a point are
associative.
Lemma (Associativity). Let (X, Fx ) be a topological space, and let x0 ∈ X. Let [f ], [g], [h] ∈
π1 (X, x0 ). Then:
([f ][g])[h] = [f ]([g][h])
Proof. Before actually proving the result, we will explicitly write the formulas for
(f ∗ g) ∗ h and f ∗ (g ∗ h):
f (4s)
s ∈ [0, 41 ] f (2s)
s ∈ [0, 21 ]
(f ∗ g) ∗ h = g(4s − 1) s ∈ [ 14 , 12 ] f ∗ (g ∗ h) = g(4s − 2) s ∈ [ 21 , 34 ]
1
h(4s − 3) s ∈ [ 43 , 1]
h(2s − 1) s ∈ [ 2 , 1]
f ½ g ¾ h
f ¼ g ½ h
70 3. DISTINGUISHING SPACES
Based on our formulas for (f ∗g)∗h and f ∗(g∗h), we parameterize the function F : I×I → X
by:
4s
s ∈ 0, 1+t
f 1+t
4
1+t 2+t
F (s, t) = g(4s − 1 − t) s ∈ 4 , 4
h 4s − 2+t
s∈
2+t
,1
2−t 2−t 4
We claim and will verify that F is a homotopy from (f ∗ g) ∗ h to f ∗ (g ∗ h). We first prove
continuity using the pasting lemma as usual. The proof that each of the piecewise parts of
F is defined on a closed set is left as a (relatively trivial but tedious) exercise.
We will, however, show that on the boundaries between these regions, the piecewise parts
agree. We see:
4 1+t
f · = f (1) = x0
1+t 4
1+t
g 4· −1−t = g(0) = x0
4
2+t
g 4· −1−t = g(1) = x0
4
4 2+t 2+t
h · − = h(0) = x0
2−t 4 2−t
so continuity follows by
the pasting
lemma. Note that we assume f, g, h are continuous
4s 2+t
so that things like h 2−t − 2−t are because compositions of continuous functions are
4s
continuous and over t ∈ I, 2−t − 2+t
2−t is continuous. The justification is similar for the other
piecewise parts in the definition of F .
We now verify that F gives a path from x0 to x0 for every fixed t ∈ I. We see that:
F (0, t) = f (0) = x0
F (1, t) = h(1) = x0
as desired.
Finally, we need to show that F provides a homotopy from (f ∗ g) ∗ h to f ∗ (g ∗ h). We
have that:
s ∈ 0, 1+t
f (4s)
4
F (s, 0) = g(4s − 1) s ∈ 1+t 2+t
4 , 4
= (f ∗ g) ∗ h
2+t
h (2s − 1) s ∈ 4 , 1
7. LOOPS AND THE FUNDAMENTAL GROUP 71
s ∈ 0, 1+t
f (2s)
4
F (s, 1) = g(4s − 2) s ∈ 1+t , 2+t
4 4
= f ∗ (g ∗ h)
h (4s − 3) s ∈ 2+t
4 ,1
Define F : I × I → X by
(
2s
s ∈ 0, t+1
f t+1 2
F (s, t) =
s ∈ t+1
ex1 2 ,1
This is:
Continuous: By the Pasting lemma, as f, ex1 are continuous on closed domains it suffices to
t+1
2
check that they agree for s = t+1
2 . This follows easily, as f
2
t+1 = f (1) = x1 ,
and ex1 t+1
2 = x1 .
A homotopy: (
f (2s) s ∈ 0, 12
F (s, 0) = = f ∗ ex1
ex1 (t) s ∈ 21 , 1
and (
f (s) s ∈ [0, 1]
F (s, 1) =
e x1 s ∈ [1, 1] = f.
We’ve now shown that π1 (X, x0 ) is closed, has an identity, and the operation is associative,
so just showing that inverses exist proves that it’s a group.
Lemma (Inverses). Let f be a path in X from x0 to x1 . Then f ∗ f ∼ ex0 and f ∗ f ∼ ex0 .
The “wrong” approach: increase the speed over f and f and wait at x1 . The “right”
approach: travel successively smaller distances along f .
Proof. We show only that f ∗ f ∼ ex0 , and the other case follows similarly.
Define F : I × I → X by
s ∈ 0, 1−t
f (2s)
1−t 2 1+t
F (s, t) = f (1 − t) s∈ 2 , 2
f (2s − 1) s ∈ 1+t
, 1
2
This is:
Continuous: Using the pasting lemma, it suffices to check that these functions agree on their
(shared) endpoints. As f (2( 1−t 1+t
2 )) = f (1 − t), and f (2( 2 ) − 1) = f (t) = f (1 − t),
we conclude that F is continuous.
A homotopy: Let t = 0,
f (2s)
s ∈ [0, 1/2]
F (s, 0) = f (1) s ∈ [1/2, 1/2] = f ∗ f (s)
f (2s − 1) s ∈ [1/2, 1]
For t = 1,
f (2s)
s ∈ [0, 0]
F (s, 1) = f (0) s ∈ [0, 1] = ex0 .
f (2s − 1) s ∈ [1, 1]
It follows that the action ∗ defines a group on the set of loops based at x0 .
Observe that the set of paths does not have a group structure, as there is no definition of
multiplication between arbitrary paths.
Example. Let X = Rn and x0 be a point in X. Then π1 (X, x0 ) = h[ex0 ]i.
The following theorem relates the fundamental group at a point within a path component
to the fundamental group of that point in the ambient space:
7. LOOPS AND THE FUNDAMENTAL GROUP 73
Theorem 25. Let A be a path component of a topological space X, and let x0 ∈ A. Then:
π1 (A, x0 ) ∼
= π1 (X, x0 )
(Note that ∼
= denotes a group isomorphism and not a homeomorphism)
Before proving the theorem, we will cover two quick non-examples of cases where the
theorem could break down.
Recall the circle S 1 which we can consider as a subspace of R2 . Since R2 is a simply
connected space, the fundamental group at every point is trivial. On the other hand,
picking some point x0 ∈ S 1 , the loop around the circle cannot be deformed in S 1 to the
point x0 , so π1 (S 1 , x0 ) is non-trivial and hence not isomorphic to π1 (R2 , x0 ) if we embed
S 1 in R2 . The following picture illustrates S 1 with the point x0 :
r0
'$
x
&%
This may seem like a counterexample to the theorem, but in R2 , S 1 is not a distinct path
component of R2 , so the theorem does not apply. Intuitively, by embedding S 1 in R2 , the
interior of the circle is part of R2 , so we can deform a loop around S 1 based at x0 to the
trivial loop by “pulling” the loop through the middle of the circle, which we could not do
when S 1 was considered as a space in its own right.
Consider the following diagram:
r0
'$
x
By A
y
&%
where we have the space R2 with A, B removed, so loops based at x0 that run around A, B
cannot be deformed into the trivial loop. However, when we consider this space embedded
in R2 as a subspace, it is not an entire path component in R2 , so our theorem does not
apply.
We now prove the theorem:
74 3. DISTINGUISHING SPACES
Before providing the proof, we will define a key map which we will use again later in the
course:
Definition. Define uf : π1 (X, x) → π1 (X, y) by uf ([g]) = [f ∗ g ∗ f ].3
We now prove the theorem by showing that uf is an isomorphism from π1 (X, x) to π1 (X, y):
Keep in mind that our overall goal is to show that the fundamental group is a topological
invariant. In general, a topological invariant is a “thing” that takes a value on topological
spaces such that every homeomorphic space takes the same value. These invariants are
used to distinguish topological spaces.
3Technically, f ∗ g ∗ f should have some indication of ordering of operations, but we dispense with these
designations because we previously proved that ∗ is associative.
76 3. DISTINGUISHING SPACES
Small Fact (about induced maps). Let φ : X → Y and φ(x0 ) = y0 . Then φ∗ is well
defined.
Proof. Let f and g be loops in X based at x0 such that f ∼ g. Then there exists
F : I × I → X, a path homotopy from f to g.
Proof. Let [f ], [g] ∈ π1 (X, x0 ). We want to show that φ∗ ([f ]X [g]X ) = φ∗ ([f ]X )φ∗ ([g]X .
Observe that
So φ ◦ f ∗ g = φ ◦ f ∗ φ ◦ g and, in particular,
φ(f ∗ g)]Y = [(φ ◦ f ) ∗ (φ ◦ g)]Y
= [φ ◦ f ]Y [φ ◦ g]Y
= φ∗ ([f ]X )φ∗ ([g]X )
This concludes the proof of the lemma.
Theorem 27. Let φ : X → Y be a homeomorphism and φ(x0 ) = y0 . Then φ∗ is an
isomorphism.
With the previous lemma we showed that φ∗ is a homomorphism. It remains to show that
φ∗ is a bijection.
Proof. 1-1: Let [f ]X , [g]X ∈ π1 (X, x0 ) such that φ∗ ([f ]X ) = φ∗ ([g]X ). Then by
definition of φ∗ , [φ ◦ f ]Y = [φ ◦ g]Y .
Observe that φ ◦ f ∼Y φ ◦ g, so there exists a path homotopy F from φ ◦ f to φ ◦ g.
Also note that, as φ is a homeomorphism, φ−1 : Y → X is continuous.
Hence φ−1 ◦ F : I × I → Y is a path homotopy from φ−1 ◦ φ ◦ f to φ−1 φ ◦ g. Since
g is bijective, this is a path homotopy from f to g.
Onto: Recall that φ∗ : π1 (X, x0 ) → π1 (Y, y0 ). Let [f ] ∈ π1 (Y, y0 ). Then [φ−1 (f )]X ∈
π1 (X, x0 ), because φ−1 is continuous.
φ∗ ([φ−1 (f )]X ) = [φ ◦ φ−1 (f )]Y , and as φ is a bijection, this is [f ]Y .
This illustrates that φ∗ is bijective, and hence is an isomorphism between X and Y .
Small Fact (about induced homomorphisms). (1) If φ : X → Y and ψ : Y → Z
are continuous, then (ψ ◦ φ)∗ = ψ∗ ◦ φ∗
(2) If i : X → X is the identity, then i∗ is the identity isomorphism
(3) Let φ : X → Y be continuous and f a path in X from p to q. Then φ∗ ◦ uf =
uφ(f ) ◦ φ∗ . (recall, uf : π1 (X, p) → π1 (X, q) by uf ([g]) = [f ∗ g ∗ f ])
Proof. (1) Note that (ψ ◦ φ)∗ : π1 (X, x0 ) → π1 (Z, (ψ ◦ φ)(x0 )), a mapping from
equivalence classes of loops in X based at x0 to the equivalence classes of loops in
Z based at (ψ ◦ φ)(x0 ). Let [f ]X ∈ π1 (X, x0 ). We then have that (ψ ◦ φ)∗ ([f ]X ) =
[(ψ ◦φ)(f )]Z . Similarly, we note that (ψ∗ ◦φ∗ )([f ]X ) = ψ∗ ([φ◦f ]Y ) = [ψ ◦φ◦f ]Z =
[(ψ ◦ φ)(f )]Z .
(2) Note that i∗ : π1 (X, x0 ) → π1 (X, x0 ). Let [f ]X ∈ π1 (X, x0 ). Then i∗ ([f ]X ) =
[i(f )]X = [f ]X .
78 3. DISTINGUISHING SPACES
uf
π1 (X, p) −−−−→ π1 (X, q)
φ φ
y ∗ y ∗
uφ(f )
π1 (Y, φ(p)) −−−−→ π1 (Y, φ(q))
Proof.
(⇒) Suppose that π1 (X, x0 ) = h[ex0 ]i. Let p, q ∈ X, and f, g paths in X from p to q. Then
f ∗ g is a loop based at p. So it is trivial to see that π1 (X, p) ∼
= π1 (X, x0 ) by an earlier
theorem, from which we can see that f ∗ g ∼ ep . Using our multiplication and inverse
lemmas for path multiplication, we conclude that f ∼ g.
(⇐) Suppose that ∀p, q ∈ X and paths f, g from p to q, we have that f ∼ g. Let
p = q = x0 , let f = ex0 , and let g be a loop in X based at x0 . Then g ∼ ex0 . Hence,
π1 (X, x0 ) = h[ex0 ]i.
7.3. A Technical Lemma. We’d like to try to prove that spaces that are homotopy
equivalent have isomorphic fundamental groups. First, we’ll need a technical lemma.
Lemma (A Technical Theorem). Let φ, ψ : X → Y be continuous, and φ ' ψ by a homotopy
F . Let x0 ∈ X, and a path f : I → Y be given by f (t) = F (x0 , t). Then uf ◦ φ∗ = ψ∗ .
7. LOOPS AND THE FUNDAMENTAL GROUP 79
Let [g] ∈ π1 (X, x0 ). We want to show that f ∗ φ(g) ∗ f ∼ ψ(g). More specifically, we will
do so by showing that f ∗ φ(g) ∗ f ∼ eψ(x0 ) ∗ ψ(g) ∗ eψ(x0 ) .
Define G : I × I → Y by
s ∈ 0, 4t
eψ(x0 )
s ∈ 4t , 14
f (4s − t)
s ∈ 41 , 12
G(s, t) = F (g(4s − 1), t)
s ∈ 12 , 2−t
f (2s − 1 + t)
2
s ∈ 2−t
e
ψ(x0 ) 2 ,1
To prove that f ∗ φ(g) ∗f ∼ eψ(x0 ) ∗ ψ(g) ∗eψ(x0 ) , we must show that G is path homotopy
from f ∗ φ(g) ∗ f to eψ(x0 ) ∗ ψ(g) ∗ eψ(x0 ) .
Continuous. G is defined differently over the five intervals. Over each interval, G is the com-
position of continuous functions and hence continuous. For G to be continuous
everywhere, the value of G at the intersection of any two intervals must agree.
– s = 41 : f (4 14 − 1) = f (1 − t) = f (t).
F (g(4 41 − 1), t) = F (g(0), t) = F (x0 , t) = f (t).
– s = 2−t 2−t
2 : f (2 2 − 1 + t) = f (1) = ψ(x0 ) = eψ(x0 ) .
By the Pasting Lemma, the function G is continuous.
Homotopy. Consider G(s, 0):
eψ(x0 ) s ∈ [0, 0]
s ∈ 0, 14
f (4s)
s ∈ 14 , 12
G(s, 0) = F (g(4s − 1), 0)
1
f (2s − 1) s ∈ 2, 1
e
ψ(x0 ) s ∈ [1, 1]
80 3. DISTINGUISHING SPACES
8. Covering Maps
Example. Let X = D2 with the usual topology, and Xe = D2 × N with the usual product
topology. Define p : X e → X by p(x, n) = x. We can take U to be any open set in X;
∞
[
p−1 (U ) = p−1 (U ) ∩ Vi , where Vi = D2 × i.
i=1
...
f -1(U)
~
X
U
X
Example (a non-example). Let Xe = S 1 , and let X = S 1 ∨ S 1 . That is, X is two copies of
S 1 , which agree at a point.
Let ∼ be the equivalence relation on S 1 given by x ∼ y if and only if x, y ∈ {x1 , x2 } or
x = y. Let p the the quotient map from X e to X, corresponding to this relation, and let U
be an open set in X containing x1 , x2 , as shown. Is U evenly covered?
The answer is no. To see why, note that p−1 (U ) = V1 ∪ V2 , where V1 , V2 are disjoint open
e But p|V1 : V1 → U is not a homeomorphism because it is not onto.
sets in X.
Definition. Let p : X e X be a continuous surjection. Suppose for all x ∈ X, there
exists an evenly covered open set U containing X. We say p is a covering map, X e is the
covering space and X is the base space.
Example (another non-example). Let X e = R2 , X = R, and P : R2 → R be defined by
p(x, y) = x. Then for each open U ⊆ X, p−1 (U ) = ∪α∈A Vα . (We can think of p−1 (U ) as a
horizontal stack of uncountably many copies of U ). For each α ∈ A, p | Vα : Vα → U is a
homeomorphism. However, each Vα is not open in X. e
Lemma (Important Lemma on Covering Maps). Let p : X e → X be a covering map. Let
−1
x ∈ X. Then the subspace topology on p ({x}) is the discrete topology.
Proof. Let y ∈ p−1 ({x}). We want to show that {y} is open in p−1 ({x}). Since p is
a covering map, there exists a evenly covered open set U containing x. Then p−1 (U ) =
82 3. DISTINGUISHING SPACES
The take-home message is that in a covering space, points in the pre-image of a single point
are “spread out.”
Example (Important). Let p : R → S 1 be defined by p(x) = (cos(2πx), sin(2πx)). (The
“slinky” space.)
0 1 2 3
Not all quotient maps are covering maps. But we will prove that all covering maps are
quotient maps. (Remember our important example?)
e → X be a covering map. Then
Theorem 28. Let p : X
(1) p is an open map
(2) X is a quotient space, and p is a quotient map.
8. COVERING MAPS 83
Proof. Let U be open in X. e We want to show that p(U ) is open. Let x ∈ p(U ). Then
there exists an evenly covered open set V containing X. Now, there exists y ∈ U such that
p(y) = x. Then p−1 (V ) = ∪α∈A Vα such that the Vα are disjoint open sets and p | Vα is a
homeomorphism for all α ∈ A. Let α0 ∈ A such that y ∈ Vα0 . There exists such an α0
because x ∈ V .
e Recall p | Vα : Vα → V is a
Because U and Vα0 are both open, U ∩ Vα0 is open in X. 0 0
homeomorphism, so
To show X has the quotient topology with regard to p, we want to show that FX =
{U ⊆ X | p−1 (U ) ∈ FX }.
(⊇) Let U ⊆ X such that p−1 (U ) ∈ FXe . Because p is open, p(p−1 (U ) is open in X.
Because p is onto p(p−1 (U )) = U . Thus, U ∈ FX .
8.1. Lifts.
~
~ X 0 1
f
p
I
0 1
f
S1
Lemma. Uniqueness of Lifts. Let p : X e → X be a covering map and f : Y → X be a
covering and f : Y → X be continuous and Y be connected. Let fe0 and fe1 be lifts of f .
Suppose there exists y0 ∈ Y such that fe0 (y0 ) = fe1 (y0 ) then fe0 = fe1 .
Let z ∈ fe0−1 (Vα1 ) ∩ fe1−1 (Vα2 ), implying fe0 (z) ∈ Vα1 and fe1 (z) ∈ Vα2 . We now want
to show that fe0 (z) 6= fe1 (z). Recall that p | Vα1 is 1 − 1 and p ◦ fe0 (y) = f (y) =
p ◦ fe1 (y): because fe0 (y) 6= fe1 (y) where fe0 (y) ∈ Vα1 and fe1 (y) ∈ Vα2 , implying that
α1 6= α2 . Otherwise, fe0 (y) and fe1 (y) would be two points in Vα1 that both map
to f (y). Therefore, Vα1 ∩ Vα2 = ∅.
Therefore fe0 (z) 6= fe1 (z), implying that z ∈ Y − Y 0 . This implies y is contained in
the open set fe0−1 (Vα1 ) ∩ fe1−1 (Vα2 ) ⊆ Y − Y 0 , making Y − Y 0 open.
Therefore Y 0 is clopen in Y . Because Y 0 is non-empty and Y is connected, Y 0 must be all
of Y . By the definition of Y 0 , fe0 = fe1 .
Lemma (Lebesgue Number Lemma). Let X be a compact metric space and let Ω be an
open cover of X. Then ∃ r > 0 such that ∀ A ⊆ X with lub{d(p, q)|p, q ∈ A} < r, A is
contained in a single element of Ω.
(r is said to be a Lebesgue Number for Ω)
fe( n1 ) ∈ Vα2 . So, as above, define fe: [ n1 , n2 ] → X e by fe(s) = (p|Vα )−1 f (s). fe:
2
2
[0, n ] → X is therefore continuous by Pasting Lemma.
e
Bi = [ i−1 i
n , n ]. ∀ i, j, F (Ai × Bj ) ⊆ Vij for some Vij ∈ {Vx }. By Part 1, we can lift
F |(I × {0} ∪ {0} × I) to Fe : (I × {0} ∪ {0} × I) → X e such that Fe(0, 0) = a ∈ X.
e
Now we worry that this choice of Vα11 will agree with how we defined Fe on the
set L. Worry not! For L is connected, and L ∩ (I1 × J1 ) is connected, and since Fe
is continuous, Fe(L ∩ (I1 × J1 )) is connected. Since the Vα are open and disjoint,
we may therefore conclude that Fe(L ∩ (I1 × J1 )) ⊆ Vα11 (otherwise it would be
disconnected).
Since V1 1 was evenly covered, we know that p | Vα11 is a homeomorphism, so we
may define Fe : I1 × J1 → X
e by:
The natural intuition is that our new function Fe is a path homotopy when F is a path
homotopy. This intuition provides a delightful segue to the next theorem:
Theorem 30 (Monodromy Theorem). Let p : X e → X be a covering map, and let a ∈ X. e
Let x1 , x2 ∈ X. Suppose that p(a) = x1 , and that f, g are paths in X from x1 to x2 . Let
fe, ge be the unique lifts of f, g beginning at a. Then if f ∼ g, fe(1) = ge(1) and fe ∼ ge.
Before beginning the proof, we observe with relish the etymology of monodromy. Mono
being the prefix for one, and dromy being some sort of Greek for a race track. E.g.
8. COVERING MAPS 87
p ◦ Fe | (I × {0}) = F | (I × {0}) = f
p ◦ Fe | (I × {1}) = F | (I × {1}) = g
p ◦ Fe | ({0} × I) = F | ({0} × I) = x1 .
Now we know that since p is a covering map, p−1 (x1 ) has the discrete topology. Also,
Fe({0} × I) is connected, so must contain only a single point of p−1 (x1 ). Certainly a ∈
Fe({0} × I), so we may say that a = Fe({0} × I) as desired.
The previous consideration tells us that Fe | (I × {1}) = ge, since the left hand side is a lift
of g originating at a, and by the uniqueness of lifts this must be ge. To finish off proving
that Fe is a path homotopy, we need to show that the endpoints are constant as well. That
is, we want to show that Fe({1} × I) = a0 for some a0 ∈ X.
e But for this, the same argument
as above applies, replacing every instance of x1 with x2 . So we conclude that:
which was part of what we were trying to prove. All these considerations together tell us
that Fe is a path homotopy between fe and ge, so fe ∼ ge and we are done.
88 3. DISTINGUISHING SPACES
8.2. S 1 and Z. Apparently our ultimate goal is to prove that the fundamental group
of S 1 is isomorphic to Z (with addition). But we need just a teensy bit more machinery,
and introduce a new function.
Definition. Let p : R → S 1 be the covering map p(x) = (cos 2πx, sin 2πx). Let x0 = (1, 0),
let f be a loop in S 1 with base point x0 . Define the degree of f , denoted deg(f ), as fe(1)
where fe is the unique lift of f starting at 0.
First, observe that deg(f ) is well defined. That is, it doesn’t matter which lift we select,
because there is only one lift! Also, convince yourself that p−1 ({x0 }) = Z.
This leads us to the theorem we have been clamoring for:
Theorem 31. Let x0 ∈ S 1 . Then π1 (S 1 , x0 ) ∼
= (Z, +).
the lift of f ∗ g beginning at 0. We want to show that f] ∗ g(1) = fe(1) + ge(1). Let
m = fe(1), n = ge(1). Let us define a function h : I → R by:
(
fe(2s) s ∈ [0, 21 ]
h(s) =
ge(2s − 1) + m s ∈ [ 12 , 1]
We rejoice at the fact that we have now seen a non trivial fundamental group.
Proof. We only needed simple connectedness in the proof that φ was 1 − 1 and onto,
so we use an analogous proof to show that φ : π1 (X, x0 ) → p−1 ({x0 }) by φ([f ]) = fe(1),
where fe is the unique lift originating at some y0 ∈ p−1 ({x0 }).
Theorem 33. Let p : S 2 → RP2 be the quotient map. Then p is a covering map.
90 3. DISTINGUISHING SPACES
Proof. Recall that the equivalence relation for this quotient map was defined as x ∼ y
iff x = ±y. We use one of the most powerful proof techniques known to mathematics here:
proof by picture. Consider any open ball in RP2 containing some point. Then its pre image
will be a pair of disjoint balls such that p restricted to the balls will be a homeomorphism.
Just think about it... or look at the pictures below.
-1 S2
2 f
RP2 provides us with another example of a space with a nontrivial fundamental group.
From one of our theorems, we know that given x0 ∈ RP2 there is a bijection from
π1 (RP2 , x0 ) to p−1 ({x0 }) (since RP2 is simply connected), and we know that p−1 ({x0 })
has precisely two elements, so π1 (RP2 , x0 ) ∼= Z2 (alternatively written by algebraists as
Z/2Z).
9. R2 ∼
6= R3
Theorem 34. R2 ∼
6= R3
Example (Not all Fundamental Groups are Abelian). Let X = S 1 ∨ S 1 with wedge point
x0 . Then π1 (X, x0 ) is not abelian.
~
X X
Now define fe(t) = (t, 0), ge(t) = (0, t), and f = p ◦ fe, g = p ◦ ge. So f is a single loop on the
first circle and g is a single loop on the second circle.
Next, lift f ∗ g and g ∗ f at the origin. The construction of X e gives that f] ∗ g(1) = (1, 0)
and g] ∗ f (1) = (0, 1). So, by the Monodromy theorem, as this are lifts with the same fixed
point, f ∗ g 6∼ g ∗ f . Therefore [f ] and [g] satisfy [f ], [g] ∈ π1 (X, x0 ) and [f ][g] 6= [g][f ], so
π1 (X, x0 ) is not abelian, as desired.