0% found this document useful (0 votes)
21 views32 pages

đồng luân và nhóm cơ bản

Chapter 7 discusses the fundamental group of a manifold, introduced by Henri Poincaré, which relates to the concept of homotopy between continuous functions. It defines the fundamental group as the set of equivalence classes of based loops, demonstrating that this set forms a group structure with properties such as associativity and the existence of inverses. The chapter also establishes that the fundamental group is isomorphic for any two base points in a path-connected space and describes how continuous mappings induce homomorphisms between fundamental groups.

Uploaded by

nguyenhoangthanh
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
21 views32 pages

đồng luân và nhóm cơ bản

Chapter 7 discusses the fundamental group of a manifold, introduced by Henri Poincaré, which relates to the concept of homotopy between continuous functions. It defines the fundamental group as the set of equivalence classes of based loops, demonstrating that this set forms a group structure with properties such as associativity and the existence of inverses. The chapter also establishes that the fundamental group is isomorphic for any two base points in a path-connected space and describes how continuous mappings induce homomorphisms between fundamental groups.

Uploaded by

nguyenhoangthanh
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 32

Chapter 7

Homotopy and the


Fundamental Group

The group G will be called the fundamental group


of the manifold V .

Henri Poincaré, 1895

The properties of a topological space that we have developed so


far have depended on the choice of topology, the collection of open
sets. Taking a different tack, we introduce a different structure, alge-
braic in nature, associated to a space together with a choice of base
point (X, x0 ). This structure will allow us to bring to bear the power
of algebraic arguments. The fundamental group was introduced by
Poincaré in his investigations of the action of a group on a manifold
[66].
The first step in defining the fundamental group is to study
more deeply the relation of homotopy between continuous functions
f0 : X → Y and f1 : X → Y . Recall that f0 is homotopic to f1 ,
denoted f0  f1 , if there is a continuous function (a homotopy)

H : X × [0, 1] → Y with H(x, 0) = f0 (x) and H(x, 1) = f1 (x).

The choice of notation anticipates an interpretation of the homotopy


—if we write H(x, t) = ft (x), then a homotopy is a deformation of

95
96 7. Homotopy and the Fundamental Group

the mapping f0 into the mapping f1 through the family of mappings


ft .

Theorem 7.1. The relation f  g is an equivalence relation on the


set Hom(X, Y ) of continuous mappings from X to Y .

Proof. Let f : X → Y be a given mapping. The homotopy H(x, t) =


f (x) is a continuous mapping H : X × [0, 1] → Y and so f  f .
If f0  f1 and H : X×[0, 1] → Y is a homotopy between f0 and f1 ,
then the mapping H  : X × [0, 1] → Y given by H  (x, t) = H(x, 1 − t)
is continuous and a homotopy between f1 and f0 , that is, f1  f0 .
Finally, for f0  f1 and f1  f2 , suppose that H1 : X × [0, 1] →
Y is a homotopy between f0 and f1 , and H2 : X × [0, 1] → Y is a
homotopy between f1 and f2 . Define the homotopy H : X ×[0, 1] → Y
by 
H1 (x, 2t), if 0 ≤ t ≤ 1/2,
H(x, t) =
H2 (x, 2t − 1), if 1/2 ≤ t ≤ 1.
Since H1 (x, 1) = f1 (x) = H2 (x, 0), the piecewise definition of H gives
a continuous function (Theorem 4.4). By definition, H(x, 0) = f0 (x)
and H(x, 1) = f2 (x) and so f0  f2 . 

We denote the equivalence class under homotopy of a mapping


f : X → Y by [f ] and the set of homotopy classes of maps between
X and Y by [X, Y ]. If F : W → X and G : Y → Z are continuous
mappings, then the sets [X, Y ], [W, X], and [Y, Z] are related.

Proposition 7.2. Continuous mappings F : W → X and G : Y → Z


induce well-defined functions F ∗ : [X, Y ] → [W, Y ] and G∗ : [X, Y ] →
[X, Z] by F ∗ ([h]) = [h ◦ F ] and G∗ ([h]) = [G ◦ h] for [h] ∈ [X, Y ].

Proof. We need to show that if h  h , then h◦F  h ◦F and G◦h 


G ◦ h . Fixing a homotopy H : X × [0, 1] → Y with H(x, 0) = h(x)
and H(x, 1) = h (x), then the desired homotopies are HF (w, t) =
H(F (w), t) and HG (x, t) = G(H(x, t)). 

To a space X we associate a space particularly rich in struc-


ture, the mapping space of paths in X, map([0, 1], X). Recall that
map([0, 1], X) is the set of continuous mappings Hom([0, 1], X) with
7. Homotopy and the Fundamental Group 97

the compact-open topology. The space map([0, 1], X) has the follow-
ing properties.
(1) X embeds into map([0, 1], X) by associating to each point
x ∈ X the constant path cx (t) = x for all t ∈ [0, 1].
(2) Given a path λ : [0, 1] → X, we can reverse the path by com-
posing with t → 1 − t. Let λ−1 (t) = λ(1 − t).
(3) Given a pair of paths λ, µ : [0, 1] → X for which λ(1) = µ(0),
we can compose paths by

λ(2t), if 0 ≤ t ≤ 1/2,
λ ∗ µ(t) =
µ(2t − 1), if 1/2 ≤ t ≤ 1.
Thus, for certain pairs of paths λ and µ, we obtain a new path λ ∗ µ ∈
map([0, 1], X).
Composition of paths is always defined when we restrict to a
certain subspace of map([0, 1], X).

Definition 7.3. Suppose X is a space and x0 ∈ X is a choice of base


point in X. The space of based loops in X, denoted Ω(X, x0 ), is
the subspace of map([0, 1], X),
Ω(X, x0 ) = {λ ∈ map([0, 1], X) | λ(0) = λ(1) = x0 }.
Composition of loops determines a binary operation ∗ : Ω(X, x0 ) ×
Ω(X, x0 ) → Ω(X, x0 ).

We restrict the notion of homotopy when applied to the space of


based loops in X in order to stay in that space during the deformation.

Definition 7.4. Given two based loops λ and µ, a loop homotopy


between them is a homotopy of paths H : [0, 1] × [0, 1] → X with
H(t, 0) = λ(t), H(t, 1) = µ(t), and H(0, s) = H(1, s) = x0 . That is,
for each s ∈ [0, 1], the path t → H(t, s) is a loop at x0 .

The relation of loop homotopy on Ω(X, x0 ) is an equivalence rela-


tion; the proof follows the proof of Theorem 7.1. We denote the set of
equivalence classes under loop homotopy by π1 (X, x0 ) = [Ω(X, x0 )],
a notation for the first of a family of such sets, to be explained later.
As it turns out, π1 (X, x0 ) enjoys some remarkable properties.
98 7. Homotopy and the Fundamental Group

Theorem 7.5. Composition of loops induces a group structure on


π1 (X, x0 ) with identity element [cx0 (t)] and inverses given by [λ]−1 =
[λ−1 ].

λ' µ' λ' µ'

H(t,s) H'(t,s) H(2t,s) H'(2t-1,s)

λ µ λ µ

Proof. We begin by showing that composition of loops induces a


well-defined binary operation on the homotopy classes of loops. Given
[λ] and [µ], we then define [λ]∗[µ] = [λ∗µ]. Suppose that [λ] = [λ ] and
[µ] = [µ ]. We must show that λ ∗ µ  λ ∗ µ . If H : [0, 1] × [0, 1] → X
is a loop homotopy between λ and λ and H  : [0, 1] × [0, 1] → X a
loop homotopy between µ and µ , then form H  : [0, 1] × [0, 1] → X
defined by

 H(2t, s), if 0 ≤ t ≤ 1/2,
H (t, s) = 
H (2t − 1, s), if 1/2 ≤ t ≤ 1.
Since H  (0, s) = H(0, s) = x0 and H  (1, s) = H  (1, s) = x0 , H  is a
loop homotopy. Also H  (t, 0) = λ ∗ µ(t) and H  (t, 1) = λ ∗ µ (t), and
the binary operation is well defined on equivalence classes of loops.
We next show that ∗ is associative. Notice that (λ ∗ µ) ∗ ν =
λ ∗ (µ ∗ ν); we only get 1/4 of the interval for λ in the first product
and 1/2 of the interval in the second product. We define the explicit
homotopy after its picture, which makes the point more clearly:

λ µ ν

λ µ ν
7. Homotopy and the Fundamental Group 99


⎪λ(4t/(1 + s)), if 0 ≤ t ≤ (s + 1)/4,


H(t, s) = µ(4t − 1 − s), if (s + 1)/4 ≤ t ≤ (s + 2)/4,
 

⎪ 4(1 − t)

⎩ν 1 − , if (s + 2)/4 ≤ t ≤ 1.
(2 − s)
The class of the constant map, e(t) = cx0 (t) = x0 , gives the
identity for π1 (X, x0 ). To see this, we show, for all λ ∈ Ω(X, x0 ),
that λ ∗ e  λ  e ∗ λ via loop homotopies. This is accomplished in
the case λ  e ∗ λ by the homotopy:

e λ

x0


x0 , if 0 ≤ t ≤ s/2,
F (t, s) =
λ((2t − s)/(2 − s)), if s/2 ≤ t ≤ 1.
The case λ  λ ∗ e is similar. Finally, inverses are constructed by
using the reverse loop λ−1 (t) = λ(1 − t). To show that λ ∗ λ−1  e
consider the homotopy:

λ λ-1



⎪ if 0 ≤ t ≤ s/2,
⎨λ(2t),
G(t, s) = λ(s), if s/2 ≤ t ≤ 1 − (s/2),


⎩λ(2 − 2t), if 1 − (s/2) ≤ t ≤ 1.
100 7. Homotopy and the Fundamental Group

The homotopy resembles the loop, moving out for a while, waiting
a little, and then shrinking back along itself. The proof that λ−1 ∗λ 
e is similar. 
Definition 7.6. The group π1 (X, x0 ) is called the fundamental
group of X at the base point x0 .

Suppose x1 is another choice of basepoint for X. If X is path-


connected, there is a path γ : [0, 1] → X with γ(0) = x0 and γ(1) =
x1 . This path induces a mapping uγ : π1 (X, x0 ) → π1 (X, x1 ) by [λ] →
[γ −1 ∗ λ ∗ γ], that is, follow γ −1 from x1 to x0 , then follow λ around
and back to x0 , then follow γ back to x1 , all giving a loop based at
x1 . Notice
uγ ([λ] ∗ [µ]) = uγ ([λ ∗ µ])
= [γ −1 ∗ λ ∗ µ ∗ γ]
= [γ −1 ∗ λ ∗ γ ∗ γ −1 ∗ µ ∗ γ]
= [γ −1 ∗ λ ∗ γ] ∗ [γ −1 ∗ µ ∗ γ] = uγ ([λ]) ∗ uγ ([µ]).
Thus uγ is a homomorphism. The mapping uγ −1 : π1 (X, x1 ) →
π1 (X, x0 ) is an inverse, since [γ ∗ (γ −1 ∗ λ ∗ γ) ∗ γ −1 ] = [λ]. Thus
π1 (X, x0 ) is isomorphic to π1 (X, x1 ) whenever x0 is joined to x1 by a
path. Though it is a bit of a lie, we write π1 (X) for a space X that
is path-connected since any choice of basepoint gives an isomorphic
group. In this case, π1 (X) denotes an isomorphism class of groups.
Following Proposition 7.2, a continuous function f : X → Y in-
duces a mapping
f∗ : π1 (X, x0 ) → π1 (Y, f (x0 )), given by f∗ ([λ]) = [f ◦ λ].
In fact, f∗ is a homomorphism of groups:
f∗ ([λ] ∗ [µ]) = f∗ ([λ ∗ µ]) = [f ◦ (λ ∗ µ)]
= [(f ◦ λ) ∗ (f ◦ µ)] = [f ◦ λ] ∗ [f ◦ µ]
= f∗ ([λ]) ∗ f∗ ([µ]).
Furthermore, when we have continuous mappings f : X → Y and
g : Y → Z, we obtain f∗ : π1 (X, x0 ) → π1 (Y, f (x0 )) and g∗ : π1 (Y, f (x0 ))
→ π1 (Z, g ◦ f (x0 )). Observe that
g∗ ◦ f∗ ([λ]) = g∗ ([f ◦ λ]) = [g ◦ f ◦ λ] = (g ◦ f )∗ ([λ]),
7. Homotopy and the Fundamental Group 101

so we have (g ◦ f )∗ = g∗ ◦ f∗ . It is evident that the identity mapping


id : X → X induces the identity homomorphism of groups π1 (X, x0 ) →
π1 (X, x0 ). We can summarize these observations by the (post-1945)
remark that π1 is a functor from pointed spaces and pointed maps to
groups and group homomorphisms. Since we are focusing on classical
notions in topology (pre-1935) and category theory was christened
later, we will not use this language in what follows. For an introduc-
tion to this framework see [53].
The behavior of the induced homomorphisms under composition
has the following consequence.
Corollary 7.7. The fundamental group is a topological invariant of
a space. That is, if f : X → Y is a homeomorphism, then the groups
π1 (X, x0 ) and π1 (Y, f (x0 )) are isomorphic.

Proof. Suppose f : X → Y has continuous inverse g : Y → X. Then


g ◦f = idX and f ◦g = idY . It follows that g∗ ◦f∗ = id and f∗ ◦g∗ = id
on π1 (X, x0 ) and π1 (Y, f (x0 )), respectively. Thus f∗ and g∗ are group
isomorphisms. 

Before we do some calculations we derive a few more formal prop-


erties of the fundamental group. In particular, what conditions imply
π1 (X) = {e}, and how does the fundamental group behave under the
formation of subspaces, products, and quotients?
Definition 7.8. A subspace A ⊂ X is a retract of X if there is a
continuous function, the retraction, r : X → A for which r(a) = a for
all a ∈ A. The subset A ⊂ X is a deformation retraction if A is
a retract of X and the composition i ◦ r : X → A → X is homotopic
to the identity on X via a homotopy that fixes A, that is, there is a
homotopy H : X × [0, 1] → X with
H(x, 0) = x, H(x, 1) = r(x), and H(a, t) = a
for all a ∈ A and all t ∈ [0, 1].
Proposition 7.9. If A ⊂ X is a retract with retraction r : X → A,
then the inclusion i : A → X induces an injective homomorphism
i∗ : π1 (A, a) → π1 (X, a) and the retraction induces a surjective homo-
morphism r∗ : π1 (X, a) → π1 (A, a).
102 7. Homotopy and the Fundamental Group

Proof. The composite r ◦ i : A → X → A is the identity mapping


on A and so the composite r∗ ◦ i∗ : π1 (A, a) → π1 (X, a) → π1 (A, a) is
the identity on π1 (A, a). If i∗ ([λ]) = i∗ ([λ ]), then [λ] = r∗ i∗ ([λ]) =
r∗ i∗ ([λ ]) = [λ ], and so the homomorphism i∗ is injective. If [λ] ∈
π1 (A, a), then r∗ (i∗ ([λ])) = [λ] and so r∗ is onto. 
Examples. Represent the Möbius band M by glueing the left and
right edges of [0, 1]×[0, 1] with a twist (Chapter 4). Let A = {[(t, 12 )] |
0 ≤ t ≤ 1} ⊂ M be the circle in the middle of the band. After the
identification, A is homeomorphic to S 1 . Define the map r : M → A
by projecting straight down or up to this line, that is, [(t, s)] → [(t, 12 )].
It is easy to see that r is continuous and r|A = idA so we have a
retract. Thus the composite r∗ ◦ i∗ : π1 (S 1 ) → π1 (M ) → π1 (S 1 ) is
the identity on π(S 1 ).

For any space X, the inclusion followed by projection


X∼= X × {0} → X × [0, 1] → X
is the identity and so X is a retract of X × [0, 1]. In fact, X is a
deformation retraction via the deformation H : X × [0, 1] × [0, 1] →
X × [0, 1] given by H(x, t, s) = (x, ts): when s = 1, H(x, t, 1) = (x, t)
and for s = 0 we have H(x, t, 0) = (x, 0).
Recall that a subset K of Rn is convex if whenever x and y are
in K, then for all t ∈ [0, 1], tx + (1 − t)y ∈ K. If K ⊂ Rn is convex,
let x0 ∈ K. Then K is a deformation retraction of the one-point
subset {x0 } by the homotopy H(x, t) = tx0 + (1 − t)x. When t = 0
we have H(x, 0) = x and when t = 1, H(x, 1) = x0 . The retraction
K → {x0 } is thus a deformation of the identity on K. Examples of
convex subsets of Rn include Rn itself, any open ball B(x, ), and the
boxes [a1 , b1 ] × · · · × [an , bn ].
7. Homotopy and the Fundamental Group 103

More generally, there is always the retract {x0 } → X → {x0 },


which leads to the trivial homomorphisms of groups {e} → π1 (X, x0 )
→ {e}. This retract is not always a deformation retract. We call
a space contractible when it is a deformation retract of one of its
points.
Deformation retracts give isomorphic fundamental groups.
Theorem 7.10. If A is a deformation retract of X, then the inclusion
i : A → X induces an isomorphism i∗ : π1 (A, a) → π1 (X, a).

Proof. From the definition of a deformation retract, the composite


i ◦ r : X → A → X is homotopic to idX via a homotopy fixing the
points in A, that is, there is a homotopy H : X × [0, 1] → X with
H(x, 0) = i ◦ r(x), H(x, 1) = x, and H(a, t) = a for all t ∈ [0, 1]. We
show that i∗ ◦ r∗ ([λ]) = [λ]. In fact we show a little more:
Lemma 7.11. If f, g : (X, x0 ) → (Y, y0 ) are continuous functions,
homotopic through basepoint preserving maps, then f∗ = g∗ : π1 (X, x0 )
→ π1 (Y, y0 ).

Proof. Suppose there is a homotopy G : X×[0, 1] → Y with G(x, 0) =


f (x), G(x, 1) = g(x), and G(x0 , t) = y0 for all t ∈ [0, 1]. Consider a
loop based at x0 , λ : [0, 1] → X, and the compositions f ◦ λ, g ◦ λ,
and G ◦ (λ × id) : [0, 1] × [0, 1] → Y :
G(λ(s), 0) = f ◦ λ(s),
G(λ(s), 1) = g ◦ λ(s),
G(λ(0), t) = G(λ(1), t) = y0 for all t ∈ [0, 1].
Thus f∗ [λ] = [f ◦ λ] = [g ◦ λ] = g∗ [λ]. Hence f∗ = g∗ : π1 (X, x0 ) →
π1 (Y, y0 ). 

A deformation retract gives a basepoint preserving homotopy be-


tween i ◦ r and idX , so we have id = i∗ ◦ r∗ : π1 (X, a) → π1 (X, a).
By Proposition 7.9, we already know i∗ is injective; i∗ is surjective
because for [λ] any class in π1 (X, a), one has [λ] = i∗ (r∗ ([λ])). 
Examples. A convex subset of Rn is a deformation retract of any
point x0 in the set. It follows from π1 ({x0 }) = {e}, that for any
convex subset K ⊂ Rn , π1 (K, x0 ) = {e}. Of course, this includes
104 7. Homotopy and the Fundamental Group

π1 (Rn , 0) = {e}. Next consider Rn − {0}. The (n − 1)-sphere S n−1 ⊂


Rn is a deformation retract of Rn − {0} as follows: Let
F : (Rn − {0}) × [0, 1] → Rn − {0}
be given by
x
F (x, t) = (1 − t)x + t .
x
Here F (x, 0) = x and F (x, 1) = x/x ∈ S n−1 . By Theorem 7.10,
π1 (Rn − {0}, x0 /x0 ) ∼
= π1 (S n−1 , x0 /x0 ).

A space X is said to be simply-connected (or 1-connected ) if


it is path-connected and π1 (X) = {e}. Any convex subset of Rn or,
more generally, any contractible space is simply-connected. Further-
more, simple connectivity is a topological property.

Theorem 7.12. Suppose X = U ∪ V , where U and V are open,


simply-connected subspaces and U ∩ V is path-connected. Then X is
simply-connected.

Proof. Choose a point x0 ∈ U ∩ V as basepoint. Let λ : [0, 1] → X


be a loop based at x0 . Since λ is continuous, {λ−1 (U ), λ−1 (V )} is an
open cover of the compact space [0, 1]. The Lebesgue Lemma gives
points 0 = t0 < t1 < t2 < · · · < tn = 1 with λ([ti−1 , ti ]) ⊂ U or V .
We can join x0 to λ(ti ) by a path γi . Define for i ≥ 1,
λi (s) = λ((ti − ti−1 )s + ti−1 ), 0 ≤ s ≤ 1,
for the path along λ joining λ(ti−1 ) to λ(ti ).

V
U
λ .x 0
γ
1

λ(t1)
7. Homotopy and the Fundamental Group 105

Then λ  λ1 ∗ λ2 ∗ · · · ∗ λn and, furthermore,


λ  (λ1 ∗ γ1−1 ) ∗ (γ1 ∗ λ2 ∗ γ2−1 ) ∗ (γ2 ∗ λ3 ∗ γ3−1 ) ∗ · · · ∗ (γn−1 ∗ λn ).
−1
Each γi ∗ λi+1 ∗ γi+1 lies in U or V . Since U and V are simply-
connected, each of these loops is homotopic to the constant map.
Thus λ  cx0 . It follows that π1 (X, x0 ) ∼
= {e}. 

Corollary 7.13. π1 (S n ) ∼
= {e} for n ≥ 2.

Proof. We can decompose S n as a union of U = {(r0 , r1 , . . . , rn ) ∈


S n | rn > −1/4} and V = {(r0 , r1 , . . . , rn ) ∈ S n | rn < 1/4}. By
stereographic projection from each pole, we can establish that U and
V are homeomorphic to an open disk in Rn , which is convex. The
intersection U ∩ V is homeomorphic to S n−1 × (−1/4, 1/4), which is
path-connected when n ≥ 2. 

Since S n−1 ⊂ Rn − {0} is a deformation retract, we have proven:

Corollary 7.14. π1 (Rn − {0}) ∼


= {e} for n ≥ 3.

In Chapter 8 we will consider the case π1 (S 1 ) in detail.


We next consider the fundamental group of a product X × Y .

Theorem 7.15. Let (X, x0 ) and (Y, y0 ) be pointed spaces. Then


π1 (X × Y, (x0 , y0 )) is isomorphic to π1 (X, x0 ) × π1 (Y, y0 ), the direct
product of these two groups.

Recall that if G and H are groups, the direct product G × H has


underlying set the cartesian product of G and H and binary operation
(g1 , h1 ) · (g2 , h2 ) = (g1 g2 , h1 h2 ).

Proof. Recall from Chapter 4 that a mapping λ : [0, 1] → X × Y is


continuous if and only if pr1 ◦λ : [0, 1] → X and pr2 ◦λ : [0, 1] → Y
are continuous. If λ is a loop at (x0 , y0 ), then pr1 ◦λ is a loop at x0
and pr2 ◦λ is a loop at y0 . We leave it to the reader to prove that
1) If λ  λ : [0, 1] → X × Y , then pri ◦λ  pri ◦λ for i = 1, 2.
2) If we take λ∗λ : [0, 1] → X ×Y , then pri ◦(λ∗λ ) = (pri ◦λ)∗
(pri ◦λ ).
106 7. Homotopy and the Fundamental Group

These facts allow us to define a homomorphism

pr1∗ × pr2∗ : π1 (X × Y, (x0 , y0 )) → π1 (X, x0 ) × π1 (Y, y0 )

by pr1∗ × pr2∗ ([λ]) = ([pr1 ◦λ], [pr2 ◦λ]). The inverse homomorphism
is given by ([λ], [µ]) → [(λ, µ)(t)], where (λ, µ)(t) = (λ(t), µ(t)). Thus
we have an isomorphism. 

We can use such results to show that certain subspaces of a


space are not deformation retracts. For example, if π1 (X, x0 ) is
a nontrivial group, then π1 (X × X, (x0 , x0 )) is not isomorphic to
π1 (X × {x0 }, (x0 , x0 )). Although X × {x0 } is a retract of X × X
via

X × {x0 } → X × X → X × {x0 },

it is not a deformation retract of X × X.


Extra structure on a space can lead to more structure on the
fundamental group. Recall (exercises of Chapter 4) that a topolog-
ical group, (G, e), is a Hausdorff topological space with basepoint
e ∈ G together with a continuous function (the group operation)
m : G × G → G, satisfying m(g, e) = m(e, g) = g for all g ∈ G, as
well as another continuous function (the inverse) inv : G → G with
m(g, inv(g)) = e = m(inv(g), g) for all g ∈ G.
Theorem 7.15 allows us to define a new binary operation on
π1 (G, e), the composite of the isomorphism of the theorem with the
homomorphism induced by m:

µ∗ : π1 (G, e) × π1 (G, e) → π1 (G × G, (e, e)) → π1 (G, e).

We denote the binary operation by µ∗ ([λ], [ν]) = [λ  ν]. On the level


of loops, this mapping is given explicitly by (λ, µ) → λ  µ, where
(λ  µ)(t) = m(λ(t), µ(t)). We next compare this binary operation
with the usual multiplication of loops for the fundamental group.

Theorem 7.16. If G is a topological group, then π1 (G, e) is an


abelian group.
7. Homotopy and the Fundamental Group 107

Proof. We first show that  and the usual multiplication ∗ on π1 (G, e)


are actually the same binary operation! We argue as follows: Repre-
sent λ ∗ µ(t) by λ  µ (t), where
 
 λ(2t), 0 ≤ t ≤ 12 ,  e, 0 ≤ t ≤ 12 ,
λ (t) = µ (t) =
2 ≤ t ≤ 1, µ(2t − 1), 12 ≤ t ≤ 1.
1
e,
Since λ(1) = e = µ(0) and m(e, µ (t)) = µ (t), m(λ (t), e) = λ (t),
we see λ ∗ µ(t) = m(λ (t), µ (t)). We next show that λ ∗ µ is loop
homotopic to λ  µ. Define two functions h1 , h2 : [0, 1] × [0, 1] → [0, 1]
by

2t/(2 − s), 0 ≤ t ≤ 1 − (s/2),
h1 (t, s) =
1, 1 − s/2 ≤ t ≤ 1,

0, 0 ≤ t ≤ s/2,
h2 (t, s) =
(2t − s)/(2 − s), s/2 ≤ t ≤ 1.

1 0

s t s t

Let F (t, s) = m(λ(h1 (t, s)), µ(h2 (t, s))). Since it is a composition
of continuous functions, F is continuous. Notice
F (t, 0) = m(λ(h1 (t, 0)), µ(h2 (t, 0))) = m(λ(t), µ(t)) = λ  µ(t)
and
F (t, 1) = m(λ(h1 (t, 1)), µ(h2 (t, 1))) = m(λ (t), µ (t)) = λ ∗ µ(t).
Thus λ ∗ µ is loop homotopic to λ  µ and we get the same binary
operation.
Given two loops λ and µ, consider the function
G : [0, 1] × [0, 1] → G, G(t, s) = m(λ(t), µ(s)).
The four corners are mapped to e and the diagonal from the lower
left to the upper right is given by λ  µ. We will take some liberties
108 7. Homotopy and the Fundamental Group

µ λ

λ µ µ λ
µ

λ#µ e λ#µ e
λ G µ

λ λ µ λ µ

λ µ

and argue with diagrams to construct a loop homotopy from λ ∗ µ to


µ ∗ λ.
Slice the square filled in by G along the diagonal and paste in a
rectangle that is simply a product of λ  µ with an interval. Put the
resulting hexagon into a square and fill in the remaining regions as
the constant map at e, the identity element of G, in the trapezoidal
regions and as λ or µ in the triangles where the path lies along the
lines joining a vertex to the opposite side.
The diagram gives a homotopy from λ ∗ µ to µ ∗ λ. It follows then
that [λ] ∗ [µ] = [µ] ∗ [λ] and so π1 (G, e) is abelian. 

Since S 1 is the topological group of unit length complex numbers,


we have proved:
Corollary 7.17. π1 (S 1 , 1) is abelian.

Exercises
1. The unit sphere in R is the set S 0 = {−1, 1}. Show that the
set of homotopy classes of basepoint preserving mappings
[(S 0 , −1), (X, x0 )] is the same set as π0 (X), the set of path
components of X.
2. Suppose that f : X → S 2 is a continuous mapping that is
not onto. Show that f is homotopic to a constant mapping.
3. If X is a space, recall that the cone on X is the quotient
space CX = X ×[0, 1]/X ×{1}. Suppose f : X → Y is a con-
tinuous function and f is homotopic to a constant mapping
Exercises 109

cy : X → Y for some y ∈ Y . Show that there is an extension


of f , fˆ: CX → Y , so that f = fˆ ◦ i, where i : X → CX is
the inclusion, i(x) = [(x, 0)].
4. Suppose that X is a path-connected space. When is it true
that for any pair of points p, q ∈ X all paths from p to q in-
duce the same isomorphism between π1 (X, p) and π1 (X, q)?
5. Prove that a disk minus two points is a deformation retract
of a figure 8 (that is, S 1 ∨ S 1 ).
6. A starlike space is a slightly weaker notion than a convex
space—in a starlike space X ⊂ Rn , there is a point x0 ∈ X
so that for any other point y ∈ X and any t ∈ [0, 1] the
point tx0 + (1 − t)y is in X. Give an example of a starlike
space that is not convex. Show that a starlike space is a
deformation retract of a point.
7. If K = α(S 1 ) ⊂ R3 is a knot, that is, a homeomorphic image
of a circle in R3 , then the complement of the knot R3 − K
has fundamental group π1 (R3 − K). In fact, this group is an
invariant of the knot in a sense that can be made precise.
Give a plausibility argument that π1 (R2 − K) = {0}. See
[69] for a thorough treatment of this important invariant of
knots.
Chapter 8

Computations and
Covering Spaces

. . . it is necessary, in order to affirm that a mani-


fold is simply-connected, to study its fundamental
group, . . .

Henri Poincaré, 1904

We have defined the fundamental group and showed that it is a


topological invariant, that is, homeomorphic spaces have isomorphic
fundamental groups. But we have yet to consider a space whose
fundamental group is nontrivial. Two familiar spaces, S 1 and RP2 ,
will provide examples.
The method of computation focuses on the properties of the map-
pings,

w : R → S1 p : S 2 → RP2
and
w(r) = cos(2πr) + i sin(2πr) = e2πir p(x) = [±x].

These mappings share certain important properties.

Definition 8.1. Let X be a space. A covering space of X is a


path-connected space X̃ and a mapping p : X̃ → X such that, for
every x ∈ X, there is an open, path-connected subset U with x ∈ U
for which each path component of p−1 (U ) is homeomorphic to U by

111
112 8. Computations and Covering Spaces

restriction of the mapping p. Such open sets are called elementary


neighborhoods.

)
( ) ( ) ( ) U

(
w-1(U)

p
p -1(U)
U

For example, if eiθ ∈ S 1 , then for 0 < < π, the open set
U = {eiα | θ − < α < θ + } in S 1 has inverse image under w given
by  
 θ θ
−1
w (U ) = − + k, + +k .
k∈Z 2π 2π 2π 2π
Since /2π < 1/2, the intervals in the union are all disjoint. Fur-
thermore, w restricted to any one of these intervals has an inverse
given by a branch of the logarithm. In the case of the quotient map
p : S 2 → RP2 , for a connected open set V ⊂ S 2 satisfying V ∩−V = ∅,
we have p(V ) open in RP2 and p−1 (p(V )) = V ∪ −V . Since the com-
ponents of p−1 (p(V )) are V and −V , it is an elementary neighbor-
hood. For any [±x] ∈ RP2 , there is such an elementary neighborhood
containing [±x] and so p : S 2 → RP2 is a covering space.
Henceforth we will assume that all spaces are path-connected and
locally path-connected to avoid pathological cases. The most useful
property of covering spaces is the ability to lift paths in X to paths
in X̃ while preserving the homotopy relation.
Lemma 8.2. Let p : X̃ → X be a covering space and let x̃0 ∈ X̃ with
p(x̃0 ) = x0 ∈ X. If λ : [0, 1] → X is any path with λ(0) = x0 , then
there exists a unique path λ̂ : [0, 1] → X̃ with λ̂(0) = x̃0 and p ◦ λ̂ = λ.
8. Computations and Covering Spaces 113

Proof. Cover X by elementary neighborhoods. If λ([0, 1]) ⊂ U for


some elementary neighborhood, then x0 ∈ U and x̃0 ∈ p−1 (U ). It
follows that x̃0 lies in some component C0 of p−1 (U ) that is home-
omorphic to U via p|C0 : C0 → U . Let (p|C0 )−1 : U → C0 denote
the inverse of this homeomorphism and let λ̂ = (p|C0 )−1 ◦ λ. Then
λ̂(0) = (p|C0 )−1 (x0 ) = x˜0 , since x˜0 is the only point in X̃ correspond-
ing to x0 in this component. Finally, p ◦ λ̂ = p ◦ (p|C0 )−1 ◦ λ = λ.
If λ([0, 1]) ⊂ U , consider the collection {λ−1 (U  ) ⊂ [0, 1] | U  ,
an elementary neighborhood}. This is a cover of [0, 1], which is a
compact metric space, and so by Lebesgue’s Lemma we can choose
0 = t0 < t1 < · · · < tn−1 < tn = 1 with each λ([ti−1 , ti ]) a subset
of some elementary neighborhood (take ti − ti−1 < δ, the Lebesgue
number). Using the argument above, lift λ on [0, t1 ]. Then take λ(t1 )
as x0 and λ̂(t1 ) as x̃0 and lift λ to [t1 , t2 ]. Continuing in this manner,
we construct λ̂ on [0, 1] with λ̂(0) = x̃0 and p ◦ λ̂ = λ.
To show that λ̂ constructed in this manner is unique, we prove a
more general result that implies uniqueness.

Lemma 8.3. Let p : X̃ −→ X be a covering space and Y a connected,


locally connected space. Given two mappings f1 , f2 : Y → X̃ with
p ◦ f1 = p ◦ f2 , then the set {y ∈ Y | f1 (y) = f2 (y)} is either empty
or all of Y .

Proof. Consider the subset of Y given by B = {y ∈ Y | f1 (y) =


f2 (y)}. We show that B is both open and closed. If y ∈ cls B,
consider x = p ◦ f1 (y) = p ◦ f2 (y) and U an elementary neighborhood
containing x. The subset (p ◦ f1 )−1 (U ) ∩ (p ◦ f2 )−1 (U ) contains y.
Because Y is locally connected, there is an open set W for which
y ∈ W ⊂ (p◦f1 )−1 (U )∩(p◦f2 )−1 (U ) with W connected. Then f1 (W )
and f2 (W ) are connected subsets of p−1 (U ) ⊂ X̃. Since W is open
and y ∈ cls B, there is a point z ∈ W with z ∈ B. Thus f1 (z) = f2 (z)
and f1 (W ) ∩ f2 (W ) = ∅; therefore, f1 (W ) and f2 (W ) must lie in
the same component of p−1 (U ). Since p ◦ f1 (y) = p ◦ f2 (y) and the
component in which we find both f1 (y) and f2 (y) is homeomorphic
to U by the restriction of p, we have f1 (y) = f2 (y). Thus y ∈ B and
B is closed.
114 8. Computations and Covering Spaces

If we let y ∈ B, the argument above shows that the sets f1 (W )


and f2 (W ) lie in the same component C0 of p−1 (U ). It follows that,
for all w ∈ W ,
f1 (w) = (p|C0 )−1 ◦ p ◦ f1 (w) = (p|C0 )−1 ◦ p ◦ f2 (w) = f2 (w)
and so W is contained in B. Thus B is open.
The only subsets of Y that are both open and closed are Y itself
and ∅ and so we have proved the lemma. 

Using Lemma 8.3, two lifts of a path λ : [0, 1] → X which begin


at the same point in X̃ must be the same lift. This is the uniqueness
part of Lemma 8.2. 

Having lifted paths in X to paths in X̃, we next lift certain ho-


motopies between paths.

Lemma 8.4. Let p : X̃ → X be a covering space and let η0 , η1 : [0, 1] →


X̃ be two paths in X̃ with η0 (0) = η1 (0) = x̃0 . If p ◦ η0 (1) = x1 =
p ◦ η1 (1) and p ◦ η0  p ◦ η1 via a homotopy that fixes the endpoints of
the paths in X, then η1  η2 in X̃ and, in particular, η0 (1) = η1 (1).

Proof. Let H : [0, 1] × [0, 1] → X be a homotopy between p ◦ η0 and


p ◦ η1 . In this case, we have, for all s, t ∈ [0, 1],
H(s, 0) = p ◦ η0 (s) H(0, t) = p(x̃0 )
and
H(s, 1) = p ◦ η1 (s) H(1, t) = p ◦ η0 (1) = p ◦ η1 (1).
Since [0, 1] × [0, 1] is a compact metric space, when we cover it by
the collection {H −1 (U ) | U , an elementary neighborhood of X}, we
can apply Lebesgue’s Lemma to get δ > 0 for which any subset of
[0, 1] × [0, 1] of diameter < δ lies in some H −1 (U ). If we subdivide
the interval [0, 1],
0 = s0 < s1 < · · · < sm−1 < sm = 1
and
0 = t0 < t1 < · · · < tn−1 < tn = 1,
so that si − si−1 < δ/2 and tj − tj−1 < δ/2, then H maps each
subrectangle [si−1 , si ] × [tj−1 , tj ] into an elementary neighborhood
for all i and j.
8. Computations and Covering Spaces 115

To construct the lifting Ĥ : [0, 1] × [0, 1] → X̃ and show it is a


homotopy between η0 and η1 , begin by lifting H on [0, s1 ] × [0, t1 ]
to X̃ by using Ĥ = (p|C11 )−1 ◦ H, where C11 is the component of
p−1 (U11 ) containing η0 (0) and H([0, s1 ]×[0, t1 ]) ⊂ U11 , an elementary
neighborhood. Having done this, extend Ĥ next to [s1 , s2 ] × [0, t1 ].
Notice that Ĥ has been defined on the line segment {s1 } × [0, t1 ]
which is connected and this determines the component of p−1 (U21 ) for
the elementary neighborhood U21 which contains H([s1 , s2 ] × [0, t1 ]).
Once the component, say C21 , is determined, extend Ĥ by Ĥ =
(p|C21 )−1 ◦ H. Continue in this manner until Ĥ is defined on [0, 1] ×
[0, t1 ]. Next, extend to [0, 1] × [t1 , t2 ] using the fact that the value of
Ĥ has been determined on each successive subrectangle along the left
and bottom edges, as a connected subset. Continue along each row
until Ĥ is defined on [0, 1]×[0, 1]. By Lemma 8.3, Ĥ is unique fulfilling
the condition Ĥ(0, 0) = η(0). Since η0 (s) is also a lift of H(s, 0), we
have that Ĥ(s, 0) = η0 (s). The condition H(0, t) = p ◦ η0 (0) implies
that Ĥ(0, t) = η0 (0), that is, the homotopy Ĥ is constant on the
subset {0} × [0, 1]. Thus, the lift Ĥ(s, 1) of the path p ◦ η1 (s) in X
begins at η0 (0) = η1 (0), and η1 (s) is also such a lift. By uniqueness,
Ĥ(s, 1) = η1 (s). Finally, H(1, t) = p◦η0 (1) = p◦η1 (1) for all t ∈ [0, 1],
Ĥ(1, t) = η0 (1), and we conclude that η0 (1) = η1 (1) since Ĥ(1, t) is
constant. 

Uniqueness of liftings of homotopies provides considerable control


over the fundamental group through a covering space, giving us a
toehold for computation.

Corollary 8.5. Suppose p : X̂ → X is a covering space: (1) If


η : [0, 1] → X̃ is a loop at x̃0 and p ◦ η is homotopic to the constant
loop cx0 for x0 = p(x̃0 ), then η  cx̃0 . (2) The induced homomor-
phism p∗ : π1 (X̃, x̃0 ) → π1 (X, x0 ) is injective. (3) For all x ∈ X, the
subsets p−1 ({x}) of X̃ have the same cardinality.

Proof. (1) One lift of cx0 is simply the constant path cx̃0 . By
Lemma 8.4, p ◦ η  p ◦ cx̃0 = cx0 implies η  cx̃0 .
(2) If p∗ ([λ]) = p∗ ([µ]), then, because p∗ is a homomorphism,
p∗ ([λ]∗[µ−1 ]) = [cx0 ], that is, p◦(λ∗µ−1 )  cx0 . By (1), λ∗µ−1  cx̃0
or λ  µ, that is, [λ] = [µ].
116 8. Computations and Covering Spaces

(3) Suppose x0 and x1 are in X and λ : [0, 1] → X is a path


joining x0 to x1 . Suppose y ∈ p−1 ({x0 }). We define a mapping
Λ : p−1 ({x0 }) → p−1 ({x1 }) by lifting λ to λy : [0, 1] → X̃ with λy (0) =
y. Define Λ(y) = λy (1). Since λy is uniquely determined by being
a lift of p ◦ λy = λ with λy (0) = y, the function Λ is well defined.
By Lemma 8.3, lifts of λ beginning at different elements in p−1 ({x0 })
must end at different points in p−1 ({x1 }) and so Λ is injective. Using
lifts of λ−1 we deduce that Λ is surjective. (Notice that a different
choice of λ might give a different one-one correspondence Λ.) 

For w : R → S 1 , w(r) = e2πir , we find that w−1 (1 + 0i) = Z ⊂ R


and so w−1 ({z}) is countably infinite for each z ∈ S 1 . For p : S 2 →
RP2 , p−1 ({[±x0 ]}) contains two elements, x0 and −x0 . In general, if
we lift a loop ω : [0, 1] → X at x0 in X, the proof of (3) of Corollary 8.5
obtains a mapping Ω : p−1 ({x0 }) → p−1 ({x0 }) by lifting the loop. By
remark (1) of the corollary, if Ω is nontrivial, then the loop ω is not
homotopic to the constant map. This observation is enough to prove
the following.

Theorem 8.6. A. π1 (S 1 ) ∼
= Z. B. π1 (RP2 ) ∼
= Z/2Z.

Proof of A. If β : [0, 1] → S 1 is any loop at 1 ∈ S 1 , then the lift


of β to β̂ : [0, 1] → R satisfies β̂(1) ∈ Z. The properties of liftings
determine a function Ξ : π1 (S 1 ) → Z given by [β] → β̂(1).
Let α : [0, 1] → S 1 be given by α(t) = (cos(2πt), sin(2πt)). Since
α = w|[0,1] , we see that one lift of α to R is just the identity and α̂(1) =
1. It follows that α is not homotopic to the constant map at 1, c1 .
Next consider αn for n ∈ Z, given by αn (t) = (cos(2πnt), sin(2πnt)).
By the same argument for α, α̂n (1) = n and so the mapping Ξ : π1 (S 1 )
→ Z is surjective. Since each αn  c1 for n = 0, the subgroup
generated by [α], isomorphic to Z, is a subgroup of π1 (S 1 ).
To finish the proof of A, we show that if β is any loop based at 1 in
S 1 , then β  αn for some n ∈ Z. Let U1 = {(x, y) ∈ S 1 | y > −1/10}
and U2 = {(x, y) ∈ S 1 | y < 1/10}. The pair β −1 (U1 ), β −1 (U2 ) is an
open cover of [0, 1] and by Lebesgue’s Lemma we can subdivide [0, 1]
as 0 = t0 < t1 < · · · < tm−1 < tm = 1 so that
i) β([ti , ti+1 ]) ⊂ U1 or β([ti , ti+1 ]) ⊂ U2 for 0 ≤ i < m.
8. Computations and Covering Spaces 117

Form the union of consecutive subintervals when both are mapped


to the same Uj , j = 1 or 2. In detail, let s0 = 0 and s1 = ti1 ,
where β([0, ti1 ]) ⊂ Uj1 for j1 one of 1 or 2 and β([ti1 , ti1 +1 ]) ⊂ Uj1 .
Let Uj2 = Uj1 and β([ti1 , ti1 +1 ]) ⊂ Uj2 . Then let s2 = ti2 , where
β([ti1 , ti2 ]) ⊂ Uj2 but β([ti2 , ti2 +1 ]) ⊂ Uj2 . Continue in this manner
to get
0 = s0 < s1 < · · · < sk−1 < sk = 1
so that
ii) β([sj−1 , sj ]) and β([sj , sj+1 ]) are not both contained in the
same Uk , for k = 1, 2.
Let βj : [0, 1] → S 1 denote the reparameterization of β|[sj ,sj+1 ] so
that β  β0 ∗ β1 ∗ · · · ∗ βk−1 , and each βj is a path in exactly one of
U1 or U2 . Furthermore, β(sj ) ∈ U1 ∩ U2 , a subspace of two compo-
nents, one of which contains 1 = e2πi0 and the other −1 = eπi . For
0 < j < m choose a path λj : [0, 1] → U1 ∩ U2 with λj (0) = β(sj ) =
βj−1 (sj ) and λj (1) = 1 or −1, depending on the component. Define
γ1 = β0 ∗ λ1 ,
γj = λ−1
j−1 ∗ βj−1 ∗ λj for 1 < j < k,

γk = λ−1
m−1 ∗ βk−1 .

By cancelling λj ∗ λ−1j , β  γ1 ∗ γ2 ∗ · · · ∗ γk . Consider the paths γ .


If γ is a closed path, it lies in U1 or U2 , which are simply-connected
and so γ  c1 or γ  c−1 . If γ is not closed, then it crosses between
the components of U1 ∩ U2 . It follows that γ  η1±1 or γ  η2±1 ,
where η1 (t) = (cos(πt), sin(πt)), the path joining 1 to −1 in U1 , and
η2 (t) = (cos(πt + π), sin(πt + π)), the path joining −1 to 1 in U2 .
Making the cancellations of the type η1 η1−1  c1 or η2 η2−1  c−1 , we
are left with three possibilities:
β  c1 , β  η1 ∗ η2 ∗ η1 ∗ η2 ∗ · · · ∗ η1 ∗ η2 , or
β  η2−1 ∗ η1−1 ∗ η2−1 ∗ · · · ∗ η2−1 ∗ η1−1 ,
after cancelling out c±1 . The ordering is determined by the fact that
β begins and ends at 1, and each γ either joins 1 to −1, joins −1 to
1, or it simply stays put. After cancellation of the paths that stay put
or products of paths that are homotopic to the constant path, we are
118 8. Computations and Covering Spaces

left with such a product in that order. Finally, w|[0,1] = α  η1 ∗ η2


and so β  αn for some n ∈ Z. 

x0. .x 0

Proof of B. Consider the model of the projective plane given by


the di-gon, a disk with each point on the boundary identified with
the point symmetric with respect to the origin. Let x0 ∈ RP2 be the
point x0 = [±(1, 0, 0)]. Let p : S 2 → RP2 denote the covering space
p(x) = [±x]. Let the loop a in RP2 denote half of the equator, and lift
a to S 2 . We get a path â from (1, 0, 0) to (−1, 0, 0) along the equator
of S 2 . By Corollary 8.5, a  cx0 . In the di-gon representation of
RP2 , a ∗ a = a2 surrounds the disk, and so a2 can be contracted to
cx0 by shrinking to the center of the disk and translating over to x0 .
It follows that π1 (RP2 ) contains Z/2Z. To finish, we need show that
any loop at x0 is homotopic to an for some n ∈ Z. Using the di-gon
we see that away from the image of the path a2 a path lies in the
contractible interior of a disk. The disk can be used to push any loop
onto a as often as it crosses between the copies of x0 . Thus we see
that any loop based at x0 is homotopic to an for some n ∈ Z and so
homotopic to a or cx0 . This implies that
π1 (RP2 ) = [a]/([a]2 = [cx0 ]) ∼
= Z/2Z.
This completes the proof of Theorem 8.6. 

Covering spaces can be developed much further. We refer the


reader to [55] or [51] for thorough treatments. Let’s turn now to
applications. We first return to the central question of the text:

Invariance of Dimension for (2, n). For n = 2, Rn and R2 are


not homeomorphic.
8. Computations and Covering Spaces 119

Proof. We assume that n ≥ 2 since the case of n = 1 is covered


in Chapter 5. If Rn ∼ = R2 , then, by composing with a translation
if needed, we can choose a homeomorphism f : Rn → R2 for which
f (0) = (0, 0). Such a mapping induces a homeomorphism Rn − {0} ∼ =
R2 − {(0, 0)}. Since S n−1 is a deformation retract of Rn − {0}, by
Theorem 7.10, π1 (Rn − {0}) ∼ = π1 (S n−1 ). For n > 2, Corollary 7.13
n−1 ∼
states that π1 (S ) = {e}, while, for n = 2, π1 (S 1 ) ∼
= Z. Since the
fundamental group is a topological invariant, it must be the case that
n = 2. 

This argument is characteristic of the power of introducing alge-


braic structures as topological invariants of spaces. Our goal in later
chapters is to generalize these ideas.
Recall the somewhat unexpected topological property introduced
in the exercises of Chapter 2: A space X has the fixed point property
(FPP) if any continuous mapping f : X → X has a fixed point, that
is, there exists a point x0 ∈ X with f (x0 ) = x0 . By the Intermediate
Value Theorem we can prove that the interval [0, 1] has the FPP: if
f : [0, 1] → [0, 1] is continuous, then define g(x) = f (x)−x : [0, 1] → R.
If f (0) = 0 and f (1) = 1, then g(0) > 0 and g(1) < 0 and so g must
take the value 0 somewhere. If g(x) = 0, then f (x) = x.
What is the generalization of the space [0, 1] to higher dimen-
sions? Candidates include [0, 1] × [0, 1] in dimension 2 or maybe the
two-disk e2 = {x ∈ R2 | x ≤ 1} = cls B(0, 1). The choice between
these two candidates is irrelevant since the fixed point property is a
topological property and they are homeomorphic. (Can you prove
it?) We generalize the fixed point property for the interval [0, 1] to
the two-disk.
Theorem 8.7 (Brouwer’s Theorem in dimension 2). The two-disk
e2 = {x ∈ R2 | x ≤ 1} ⊂ R2 has the fixed point property.

Proof. Suppose f : e2 → e2 is a continuous function without a fixed


point. Then for each x ∈ e2 , f (x) = x. Define g : e2 → S 1 by
g(x) = intersection of the ray from f (x) to x with S 1 .

To see that g(x) is continuous on e2 , we apply some vector ge-


ometry: write Q = f (x), Z = g(x). Let O = (0, 0) and define
120 8. Computations and Covering Spaces

g(x)
x X

f(x)
O

X = (x − Q)/x − Q. Then, g(x) = Z = Q + tX for some t ≥ 0 for


which Q + tX ∈ S 1 , that is, (Q + tX) · (Q + tX) = 1. This condition
can be rewritten to solve for t, namely,
(Q + tX) · (Q + tX) = t2 (X · X) + 2t(Q · X) + Q · Q = 1.
The quadratic formula gives

tx = −Q · X + (Q · X)2 + 1 − Q · Q
 2
x−f (x) x−f (x)
= −f (x) · + f (x) · + 1−f (x) · f (x).
x−f (x) x−f (x)
Note that this choice of signs gives tx ≥ 0, and tx = 0 implies
f (x) = x. Since we have assumed that this doesn’t happen, tx > 0.
Furthermore, tx is a continuous function of x. We can write g(x)
explicitly as
x − f (x)
g(x) = f (x) + tx
x − f (x)
and so g(x) is continuous.
By the definition of the mapping g, if x ∈ S 1 ⊂ e2 , then g(x) = x.
We have constructed a continuous mapping g : e2 → S 1 for which
g ◦ i = idS 1 , that is, the identity mapping on S 1 can be factored:
i g
idS 1 : S 1 −→ e2 −→ S 1 .
This composite leads to a composite of group homomorphisms and
fundamental groups:
i ∗ g∗
id : π1 (S 1 ) −→ π1 (e2 ) −→ π1 (S 1 ).
8. Computations and Covering Spaces 121

However, π1 (e2 ) = {[c1 ]} and so g∗ ◦ i∗ ([α]) = [c1 ] = [α] and g∗ ◦ i∗ =


id, a contradiction. Therefore, a continuous function f : e2 → e2
without fixed points is not possible. 

Corollary 8.8. S 1 is not a retract of e2 .

More powerful tools will be developed in later chapters to prove a


generalization of Theorem 8.7 and its corollary. Brouwer proved this
general result around 1911 [11].
We next apply the fact that π1 (RP2 ) ∼= Z/2Z. Recall that RP2
is the space of lines through the origin in R3 . The lower dimensional
analogue is the space RP1 consisting of lines through the origin in
R2 . We can identify a line with the angle it makes with the x-axis.
To obtain every line through the origin, we only need angles 0 ≤
θ ≤ π, where the x-axis is identified with the angles 0 and π. Hence
RP1 ∼ = [0, π]/(0 ∼ π) ∼ = S 1 . Thus π1 (RP1 ) ∼ = Z. The analogue of
the covering map p : S 2 → RP2 in this case is p : S 1 → RP1 given by
e2πiθ → [±e2πiθ ]. In fact, p∗ : π1 (S 1 ) → π1 (RP1 ) is described as a
homomorphism Z → Z given by multiplication by two, because the
generator [α] wraps around RP1 twice.
In Chapter 5 we proved that a continuous mapping f : S 1 → R
must send some point and its negative to the same value, that is, there
is always a point x0 ∈ S 1 with f (x0 ) = f (−x0 ). We can generalize
that result to S 2 .

Theorem 8.9. If f : S 2 → R2 is a continuous function, then there


exists a point x ∈ S 2 with f (x) = f (−x).

We proceed by proving an associated result.

Proposition 8.10 (The Borsuk-Ulam Theorem for n = 2). There


does not exist a continuous function f : S 2 → S 1 that satisfies f (−x)
= −f (x) for all x ∈ S 2 .

Proof of the Borsuk-Ulam Theorem. Assume such a function


exists. The condition satisfied by f can be written f (±x) = ±f (x).
122 8. Computations and Covering Spaces

It follows that f induces fˆ: RP2 → RP1 and fˆ fits into a diagram:

S2
f
/ S1

p p
 
RP 2 / RP 1

for which p ◦ f = fˆ ◦ p. Consider the induced homomorphism fˆ∗ :


π1 (RP2 ) → π1 (RP1 ). By Theorem 8.6, fˆ∗ is a homomorphism Z/2Z →
Z. However, any such homomorphism must be the trivial homomor-
phism. (Do you know why?) Let λ : [0, 1] → S 2 denote a path from
the North Pole to the South Pole along a meridian of constant longi-
tude. It follows that [p ◦ λ] = [α], a generator for Z/2Z ∼ = π1 (RP2 ).
Since the North and South Poles are antipodal, these points are iden-
tified in RP1 after passage through f and p̄. Hence [p̄ ◦ f ◦ λ] is
nontrivial in π1 (RP1 ). But [p̄ ◦ f ◦ λ] = [fˆ ◦ p ◦ λ] = fˆ∗ ([p ◦ λ]) = 0, a
contradiction. 

Corollary 8.11. If f : S 2 → R2 is a continuous function such that


f (−x) = −f (x) for all x ∈ S 2 , then f (x) = (0, 0) for some x ∈ S 2 .

Proof. If not, then g(x) = f (x)/f (x) would be a continuous func-


tion g : S 2 → S 1 with g(−x) = −g(x) for all x ∈ S 2 . 

Proof of Theorem 8.9. Suppose for every x ∈ S 2 that f (x) =


f (−x). Then define g(x) = f (x)−f (−x). Notice that g is continuous,
g(−x) = −g(x), and g(x) = 0 for all x ∈ S 2 , a contradiction. 

Corollary 8.12. No subset of R2 is homeomorphic to S 2 .

The corollary tells us that there is no cartographic map homeo-


morphic to the entire sphere.
Finally, we derive an unexpected corollary of our analysis of the
fundamental group of the circle, namely, the Fundamental Theorem
of Algebra. This topological proof gives a complete proof avoiding
the difficulties in the approach of Gauss in Chapter 5 based on con-
nectedness.
8. Computations and Covering Spaces 123

The Fundamental Theorem of Algebra. If p(z) = z n +an−1 z n−1


+ · · · + a1 z + a0 is a polynomial with complex coefficients, then there
is a complex number z0 with p(z0 ) = 0.

Proof. Recall that C ∼ = R2 and the nth power mapping h : z → z n


induces a mapping h : S 1 → S 1 which can be written as eiθ → einθ .
Lifting this mapping to the covering space w : R → S 1 , it represents
n ∈ Z ∼ = π1 (S 1 ) via the identification of π1 (S 1 ) with Z given by
[β] → β̂(1).
Viewed as a mapping, h : S 1 → S 1 , h induces the homomorphism
h∗ : π1 (S 1 ) → π1 (S 1 ). The law of exponents implies that

h∗ (θ → eπimθ ) = (θ → (eπimθ )n = eπinmθ ),

that is, h∗ is multiplication by n.


We first consider a special case of the theorem—suppose

|an−1 | + |an−2 | + · · · + |a0 | < 1.

Suppose p(z) has no root in e2 = {z ∈ C | |z| ≤ 1}. Define the


mapping p̂ : e2 → R2 − {0} by p̂(z) = p(z). Restricting to S 1 = ∂e2
we get p̂| : S 1 → R2 − {0}. Since p̂| can be extended to e2 , it follows
(exercise) that p̂| is homotopic to a constant map. However, consider
the mapping

F (z, t) = z n + t(an−1 z n−1 + · · · + a0 ),

which gives a homotopy between F (z, 0) = z n and F (z, 1) = p(z).


If F (z, t) never vanishes on S 1 , the homotopy implies p̂|  z n . To
establish this condition, for |z| = 1 we estimate

|F (z, t)| ≥ |z n | − |t(an−1 z n−1 + · · · + a0 )|


≥ 1 − t(|an−1 z n−1 | + · · · + |a0 |)
= 1 − t(|an−1 | + · · · + |a0 |) > 0.

As a class in π1 (S 1 ), [(z → z n )] is not homotopic to the constant map


while p̂| is, so we get a contradiction.
124 8. Computations and Covering Spaces

To reduce the general case to this special case, let t ∈ R, t = 0,


and let u = tz. So

p(u) = un + an−1 un−1 + · · · + a1 u + a0


= (tz)n + an−1 (tz)n−1 + · · · + a1 tz + a0 .

If p(u) = 0, then
an−1 n−1 a1 a0
zn + z + · · · + n−1 z + n = 0.
t t t
So given a zero for p(u) we get a zero for p̃t (z) with p̃t (z) = z n +
an−1 n−1
t z + · · · + atn0 and vice versa. Taking t large enough we can
guarantee
a   a  a 
 n−1   1   0
  + · · · +  n−1  +  n  < 1
t t t
and we can apply the special case. 

In Chapter 7 we proved that a subspace A of a space X, which is


a deformation retract of X, shares the same fundamental group as X.
Furthermore, if X and Y are homeomorphic spaces, they share the
same fundamental group. We generalize these conditions to identify
an important relation between spaces.

Definition 8.13. Two spaces are homotopy equivalent, denoted


X  Y , if there are mappings f : X → Y and g : Y → X with
g ◦ f  idX and f ◦ g  idY .

If A ⊂ X is a deformation retract, then there is a mapping


r : X → A for which idA = r ◦ i : A → A and idX  i ◦ r : X → X.
Thus A is homotopy equivalent to X and homotopy equivalence gen-
eralizes the relation of deformation retraction. Contractible spaces
are homotopy equivalent to a one-point space so homotopy equiva-
lence is a weaker notion than homeomorphism.

Proposition 8.14. In a set of topological spaces, homotopy equiva-


lence is an equivalence relation.

Proof. It suffices to check transitivity since the other properties are


clear. Suppose X  Y and Y  Z via mappings f : X → Y , g : Y →
8. Computations and Covering Spaces 125

X, t : Y → Z, and u : Z → Y . Consider t ◦ f : X → Z and g ◦ u : Z →


X. Then
(g ◦ u) ◦ (t ◦ f )  g ◦ (u ◦ t) ◦ f  g ◦ idY ◦f = g ◦ f  idX and
(t ◦ f ) ◦ (g ◦ u)  t ◦ (f ◦ g) ◦ u  t ◦ idX ◦u = t ◦ u  idZ .

Fixing a universe, that is, a set in which all relevant spaces are
elements, the equivalence class of a space X is called its homo-
topy type. The effectiveness of the fundamental group to distinguish
spaces is limited by homotopy equivalence.
Proposition 8.15. If X and Y are homotopy-equivalent spaces via
mappings f : X → Y and g : Y → X, then the induced mappings
f∗ : π1 (X, x0 ) → π1 (Y, f (x0 )) and g∗ : π1 (Y, y0 ) → π1 (X, g(y0 )) are
isomorphisms.

Proof. Let H : X × [0, 1] → X be a homotopy between g ◦ f and idX .


Let γ : [0, 1] → X be the path γ(t) = H(x0 , t), so that γ(0) = g ◦f (x0 )
and γ(1) = x0 . We can write the induced homomorphisms:
f∗ g∗ uγ
π1 (X, x0 ) −→ π1 (Y, f (x0 )) −→ π1 (X, g ◦ f (x0 )) −→ π1 (X, x0 ).
We claim that this composite is the identity homomorphism. Consider
[λ] ∈ π1 (X, x0 ). The result of the composite on this element is
[λ] → [f ◦ λ] → [g ◦ f ◦ λ] → [γ −1 ∗ (g ◦ f ◦ λ) ∗ γ].
Apply the homotopy H to get a homotopy from g ◦ f ◦ λ to λ by
H(λ(t), s). We use this homotopy to construct one from γ −1 ∗
(g ◦ f ◦ λ) ∗ γ to λ by reparameterizing according to the diagram:
λ

H(λ(t),s)
γ γ

γ gfλ γ
126 8. Computations and Covering Spaces

In the triangles, we have taken γ and opened it into a triangle


with the pictured curves given by isobars (constant paths). It follows
from the homotopy that [γ −1 ∗ (g ◦ f ◦ λ) ∗ γ] = [λ]. This implies
that f∗ is injective and g∗ surjective. To finish the proof consider the
composite
g∗
π1 (Y, f (x0 )) −→ π1 (X, g ◦ f (x0 ))
f∗ uη
−→ π1 (Y, f ◦ g ◦ f (x0 )) −→ π1 (Y, f (x0 )),
where η : [0, 1] → Y is the path η(t) = H̄(f (x0 ), t) in the homotopy
H̄ between f ◦ g and idY . The same argument applies mutatis mu-
tandis to show that f∗ is surjective and g∗ is injective and hence both
homomorphisms are isomorphisms. 

Homotopy equivalence is cruder than homeomorphism but in-


cludes it as a special case. To give an idea of how crude homotopy
equivalence is, notice that, for all n, Rn is homotopy equivalent to
a point. The letters of the alphabet as subspaces of R2 show other
failures to distinguish between different topological spaces.
A  D  S 1 , B  S 1 ∨ S 1 , C  E  F  ∗, . . . .

Proposition 8.15 shows that the fundamental group is a homo-


topy invariant, that is, if X  Y , then π1 (X) ∼= π1 (Y ). Thinking
of the fundamental group as a filter that distinguishes spaces, it can
only hope to catch homotopy inequivalent spaces. In later chapters
we will consider other homotopy invariants. Poincaré [66] introduced
the fundamental group to distinguish certain manifolds that were in-
distinguishable via other combinatorial invariants.

Exercises
1. Suppose that f : S 1 → S 1 has an extension fˆ: e2 → S 1 ,
that is, the mapping fˆ satisfies fˆ ◦ i = f , where i : S 1 → e2
is the inclusion. Show that f is null-homotopic, that is, f
is homotopic to the constant mapping.
2. Though we will not prove it, one of the useful theorems for
computing the fundamental groups of spaces is the Seifert-
vanKampen Theorem. A special case of this theorem
Exercises 127

is the following: If a path-connected space X is a union


X = U ∪ V with V simply-connected and x0 ∈ U ∩ V , then
the inclusion i : U → X induces a surjection i∗ : π1 (U, x0 ) →
π1 (X, x0 ) with kernel given by the smallest subgroup of
π1 (U, x0 ) containing j∗ (π1 (U ∩ V, x0 )), where j : U ∩ V → U
denotes the inclusion. Use the descriptions of RP2 of previ-
ous chapters and this theorem to make another computation
of π1 (RP2 ).
3. Suppose that X is simply-connected and p : X̃ → X is a
covering space of X. Show that p is a homeomorphism.
4. Let Ω(X, x0 ) denote the based loop space of X given by
Ω(X, x0 ) = {λ : [0, 1] → X | λ is continuous and λ(0) = λ(1) = x0 }.
This subspace of map(I, X) is topologized with the compact-
open topology. Show that
i) π0 (Ω(X, x0 )), the collection of path-components of
Ω(X, x0 ), is in one-one correspondence with π1 (X, x0 ).
ii) Show that the loop multiplication m : Ω(X, x0 ) ×
Ω(X, x0 ) → Ω(X, x0 ) given by m(λ, µ) = λ ∗ µ is a
continuous multiplication on Ω(X, x0 ).
5. We know from Theorem 7.15 and Theorem 8.6 that the fun-
damental group of the torus S 1 × S 1 is Z × Z. Use the
argument for the computation of π1 (RP2 ) ∼
= Z/2Z to prove
1 ∼
π1 (S × S ) = Z × Z by viewing the torus as a quotient of
1

[0, 1] × [0, 1].


6. Let’s make a space—take two distinct 2-spheres, S 2 , and join
them by a line segment—a space resembling dumbbells, but
with a very thin connector. Denote this space by X and
show that it is simply-connected.

You might also like