• 沒有找到結果。

Let β(A, U ) be the family of continuous functions f : K → X such that f (A

N/A
N/A
Protected

Academic year: 2022

Share "Let β(A, U ) be the family of continuous functions f : K → X such that f (A"

Copied!
10
0
0

加載中.... (立即查看全文)

全文

(1)

1. Loop Space and Higher homotopy groups

Let K be a compact Hausdorff space and X be a topological space. The space of all continuous functions from K to X is denoted by C(K, X). Let A be a closed subset of K (and thus A is compact) and U be an open subset of X. Let β(A, U ) be the family of continuous functions f : K → X such that f (A) ⊂ U, i.e.

β(A, U ) = {f ∈ C(K, X) : f (A) ⊂ U }.

The family {β(A, U )} forms a subbase of a topology on C(K, X); this topology is called the compact- open topology on C(K, X).

A pointed topological space is a topological space X together with a point x0 ∈ X. A pointed topological space is denoted by (X, x0). A morphism/map from (X, x0) to (Y, y0) is a continuous map f : X → Y such that f (x0) = y0.

A loop in X with based point x0 is a continuous map c : [0, 1] → X

such that c(0) = c(1) = x0. We may identify c with a morphism/map c : (S1, 1) → (X, x0) between pointed topological spaces. The constant loop at x0 is the constant map ex0 : (S1, 1) → (X, x0) so that ex0(t) = x0 for all t ∈ S1. The space of loops in (X, x0) has a subspace topology induced from C(S1, X). Together with the constant loop ex0, the space of loops is denoted by Ω(X, x0) and called the loop space of X with base point x0. The fundamental group of (X, x0) is defined to be

π1(X, x0) = π0(Ω(X, x0)).

Here π0(Y ) denotes the set of connected components of a topological space Y. We call π1(X, x0) the fundamental group of the pointed space (X, x0). Since Ω(X, x0) is again a pointed space, its fundamental group is denoted by π2(X, x0), i.e. π2(X, x0) = π1(Ω(X, x0)). Inductively, we define the higher homotopy groups of (X, x0) by

πn(X, x0) = πn−1(Ω(X, x0)) for n ≥ 1.

One can show that the construction of higher homotopy groups is functorial. In other words, for each n ≥ 1, πn defined a functor from the category of pointed topological spaces into the category of groups.

The following theorem is the most useful tool in homotopy theory.

Theorem 1.1. Let p : (E, e0) → (B, b0) be a morphism of pointed spaces such that p : E → B is a fiber bundle (or a fiber map). Let F = π−1(b0). For each n ≥ 1, there exists a group homomorphism δn: πn(B, b0) → πn−1(F, x0) such that the following sequence is exact:

(1.1) · · · −−−−→ πn(F, x0) −−−−→ ππn(i) n(E, e0) −−−−→ ππn(p) n(B, b0) −−−−→ πδn n−1(F, x0) −−−−→ · · · . Here i : F → X is the inclusion map.

Example 1.1. Let p : (R, 0) → (S1, 1) be the morphism of pointed spaces corresponding to the map p(t) = e2πit for t ∈ R. Then the fiber of p over 1 is Z. Study the long exact sequence of homotopy groups induced from the fiber map

Z → R → S1.

Example 1.2. Let Sn be the n-dimensional unit sphere in Rn+1, i.e.

Sn+1= {(x0, · · · , xn) ∈ Rn+1: x20+ · · · + x2n= 1}.

The antipodal map A is a continuous map Sn → Sn sending x to −x. Identify Z2 = Z/2Z with the subgroup {I, A} of the automorphism group of Sn. The quotient space Sn/Z2denoted by RPn called the n-dimensional real projective space. Study the long exact sequence of the homotopy groups induced from the fiber map

Z2→ Sn→ RPn.

1

(2)

Example 1.3. Let S2n+1 be the 2n + 1 dimensional unit sphere in Cn i.e.

S2n+1 = {z = (z1, · · · , zn) ∈ Cn : |z1|2+ · · · + |zn|2= 1}.

The one dimensional compact torus S1= {z ∈ C : |z| = 1} acts on S2n+1 by S1× S2n+17→ S2n+1, (λ, z) → λz.

The quotient space S2n+1/S1denoted by CPn is called the n-dimensional complex projective space.

Study the long exact sequence of homotopy groups induced from the fiber map S1→ S2n+1 → CPn.

(3)

2. Singular homology The standard q-simplex is the subset of Rq+1 defined by

q = {(t0, t1, · · · , tq) :

q

X

i=0

ti= 1, ti≥ 0}.

It is equipped with the subspace topology induced from the Euclidean topology on Rq+1. Let X be a topological space. A singular q-simplex in X is a continuous map

σ : ∆q→ X.

The space of singular q-simplices is denoted by

Sq = {σ : ∆q → X : σ is continuous}.

A singular q-chain over a ring R1is a function

α : Sq → R

so that {σ ∈ Sq : α(σ) 6= 0R} is a finite set. Here 0Ris the additive identity of the ring R. The space of singular q-chains in X is denoted either by Sq(X, R) or by Cq(X, R). We define an R-module structure on Cq(X, R) by

(α + β)(σ) = α(σ) + β(σ), (aα)(σ) = aα(σ), σ ∈ F ,

where α, β ∈ Cq(X, R) and a ∈ R. The 0 element of Cq(X, R) is the function 0 : Sq→ R defined by 0(σ) = 0R.

When q ∈ Z with q < 0, we set Sq(X, R) = 0 the trivial R-module.

For each σ : ∆q → X, we define a q-chain δσ: Sq → R by δσ(τ ) = δστ1R. Remark. We denote α(σ) by ασ.

Lemma 2.1. The set {δσ : σ ∈ Sq} forms an R-basis for Cq(X, R) for q ≥ 0.

Proof. Show that {δσ : σ ∈ Sq} is linearly independent over R and that α = P

σ∈Sqασδσ for any

α ∈ Cq(X, R). 

Lemma 2.2. Show that the map

δ : Sq → Cq(X, R), σ 7→ δσ

is injective.

We identify σ with its image δσ in Cq(X, R). An element α ∈ Cq(X, R) can be rewritten as α = X

σ∈F

ασσ.

2

For each 0 ≤ i ≤ q, we define a map Fqi: Rq → Rq+1by

Fqi(x0, x1, · · · , xq−1) = (x0, x1, · · · , xi−1, 0, xi, · · · , xq−1)

for (x0, x1, · · · , xq) ∈ Rq+1. Then Fqi are continuous for 0 ≤ i ≤ q. For each 0 ≤ i ≤ q, we define the i-th face map fqi : ∆q−1→ ∆q to be the restriction of Fqi to the standard q − 1 simplex fqi = Fqi|q−1. The continuities of the family of functions {fqi : 0 ≤ i ≤ q} follow from the fact that Fqi are continuous.

The i-th face of a singular q-simplex σ on X is a singular q − 1-simplex σ(i): ∆q−1→ X

1R is assumed to be commutative with multiplicative identity 1R. 2Thus 0 element in Cq(X, R) can be represented as 0 =P

σ∈F0Rσ.

(4)

defined by σ(i)= σ ◦ fqi.

The boundary of a singular q-simplex σ on X is a q − 1-chain ∂qσ on X defined by

qσ =

q

X

i=0

(−1)iσ(i).

In general, the boundary of a singular q-chain α is defined to be the q − 1-chain

qα = X

σ∈Sq

ασ(∂qσ).

When q < 0, we set ∂q= 0 to be the zero map.

Lemma 2.3. The maps ∂q : Cq(X, R) → Cq−1(X, R) is R-linear for all q ∈ Z such that

q◦ ∂q+1= 0.

Let Zq(X, R) = ker ∂q and Bq(X, R) = Im ∂q+1. Elements of Zq(X, R) are called q-cycles and elements of Bq(X, R) are called q-boundaries. Since ∂q◦ ∂q+1= 0, Bq(X, R) is an R-submodule of Zq(X, R). The quotient R-module

Hq(X, R) = Zq(X, R)/Bq(X, R)

is called the q-th singular homological module of X with coefficients in R. We set H(X, R) =M

q∈Z

Hq(X, R) called the singular homology theory of X over R. We also denote

C(X, R) =M

q∈Z

Cq(X, R).

Example 2.1. Let X = {x} be a topological space with a single point. Then Hq(X, R) =

(R if q = 0 0 otherwise.

Definition 2.1. A complex over a ring R is a sequence of R-modules {Cn: n ∈ Z} together with a sequence of R-linear maps ∂n : Cn → Cn−1 such that ∂n◦ ∂n+1= 0. A complex over R is denoted by C= (Cn, ∂n). The n-th homology of Cis defined to be

Hn(C) = ker ∂n/ Im ∂n+1.

Using the same terminology, elements of Zn(C) = ker ∂n are called n-cycles while elements of Bn(C) = Im ∂n+1are called n-boundaries.

(5)

3. Reduced Homology

Theorem 3.1. Let X be a space and {Xα: α ∈ Λ} be the set of all path components of X. For all q ≥ 0,

Hq(X, R) ∼=M

α∈Λ

Hq(Xα, R).

Hence to study the singular homology of a space, we only need to find out the singular homology of its path components. Assume that X is path connected.

The standard 0-simplex is the set ∆0 = {0}. A singular 0-simplex in X is a continuous map σ : ∆0→ X. Then σ is determined by σ(0). The set S0 and X are in one-to-one correspondence:

ι : S0→ X, σ 7→ σ(0).

We identify C0(X, R) with the space of functions α : X → R so that such that {x ∈ X : α(x) 6= 0}

is a finite set. Thus an element of C0(X, R) is represented asP

x∈Xrxx where rx∈ R.

The standard 1-simplex ∆1 is the subset {(t0, t1) ∈ R2 : t0, t1 ≥ 0, t0+ t1 = 1}. A singular simplex in X is a continuous map σ : ∆1→ X. Let h : [0, 1] → ∆1be the map t 7→ (1 − t, t). Then h is a homeomorphism3. We obtain a path σ ◦ h : [0, 1] → X in X. The space of path in X is denoted by P = C([0, 1], X). There is a bijection

S1→ P, σ 7→ σ ◦ h.

We identify C1(X, R) with the space of functions β : P → R so that {c : β(c) 6= 0} is a finite set.

We represent β as a finite sum β =P

c∈Prcc. Here rc ∈ R and c : [0, 1] → X denotes a path. Now we would like to express the boundary map ∂1using the new expression.

The face map f10: ∆0→ ∆1is the map f10(0) = (0, 1) and f11: ∆0→ ∆1is the map f11(0) = (1, 0).

If σ : ∆1→ X is a singular 1-simplex, then

σ(0)(0) = σ(0, 1) = σ ◦ h(1), σ(1)(0) = σ(1, 0) = σ ◦ h(0).

Recall the boundary of an 1-singular simplex σ in X is defined to be the 0-chain

1σ = σ(0)− σ(1).

Thus if c : [0, 1] → X is a path corresponding to the singular 1-simplex c ◦ h−1 : ∆1→ X, then by the identification described above, we find ∂1c = c(1) − c(0). If β =P

c∈Prcc, then

1β =X

c∈P

rc(c(1) − c(0))

=X

c

rcc(1) −X

c

rcc(0)

= X

x:c(1)=x

rcx − X

c:c(0)=x

rcx.

Let  : C0(X, R) → R be the map

 X

x∈X

rxx

!

= X

x∈X

rx.

Then  is surjective and thus by the first isomorphism, we find C0(X, R)/ ker  ∼= R. Observe that



 X

x:c(1)=x

rcx

=X

c

rc= 

 X

x:c(0)=x

rcx

. Since  is R-linear, we find

 ◦ ∂1β =X

c

rc−X

c

rc= 0.

3Check as an exercise

(6)

In other words,  ◦ ∂1= 0 which gives Im ∂1⊆ ker . In fact, we can prove Lemma 3.1. ker  = Im ∂1.

Proof. Let z ∈ ker . Then z = 0. Write z =P

x∈Xrxx. ThenP

x∈Xrx= 0. Denote {x ∈ X : rx6=

0} = {x1, · · · , xn} and rxi = ri for 1 ≤ i ≤ n. Then z = Pn

i=1rix. Choose x0 ∈ X. By the path connectedness of X, for each 1 ≤ i ≤ n, we choose a path βi from x0 to xn. Write β =Pn

i=1riβi. Then

∂β =

n

X

i=1

ri(xi− x0)

=

n

X

i=1

rixi

n

X

i=1

rix0

=

n

X

i=1

rixi

n

X

i=1

ri

! x0

= z

Here use the fact that r1+ · · · + rn = 0. We find z = ∂β for some β ∈ C1(X, R). We prove our

assertion. 

Since the zeroth singular homology group H0(X, R) is defined to be C0(X, R)/ Im ∂1 and ∂1 = ker , we find

H0(X, R) = C0(X, R)/ ker  ∼= R when X is path connected. In general by Theorem 3.1, we obtain:

Corollary 3.1. Let X be a space and C be the set of all path components of X. Then H0(X, R) ∼=M

C

R.

We have set Ci(X, R) = 0 when i < 0 and ∂i: Ci(X, R) → Ci−1(X, R) to be the zero map for all i < 0. In fact, we can define a new complex C0= (Ci0, ∂i0) by setting

Ci0=





Ci(X, R) if i ≥ 0

R if i = −1

0 if i < −1.

and ∂i0=





i if i > 0

 if i = 0 0 if i < 0.

.

The i-th homology of the new complex C0is called the reduced i-th homology of the space X and denoted by

Hi#(X, R) = Hi(C0).

It follows from the definition that Hi#(X, R) =

(Hi(X, R) if i 6= 0

0 if i = 0 and if X is path connected.

When X is an one pointed space, all of its reduced homology modules vanish, i.e.

Hi#(pt, R) = 0, i ∈ Z.

This motivates the definition of acyclic complex.

Definition 3.1. A complex C = (Ci, ∂i) is acyclic if all of its homological modules vanish, i.e.

Hi(C) = 0 for all i ∈ Z.

(7)

Notice that Hi(C) = ker ∂i/ Im ∂i+1 and hence C is acyclic if and only if ker ∂i= Im ∂i+1, i.e. the complex C forms an exact sequence of R-modules.

Let C = (Ci, ∂i) be a complex with the property that Ci = 0 for all i < 0. An augmentation of such a complex C over R is an R-epimorphism  : C0→ R such that ∂1 = 0. A complex with an augmentation gives us a new complex C0 defined by

Ci0 =





Ci if i ≥ 0 R if i = −1 0 if i < −1.

and ∂0i=





i if i > 0

 if i = 0 0 if i < 0.

.

The new complex C0 is called the reduced chain complex associated with (C, ). The complex C0 depends on the choice of . The corresponding homology of C0 is called the reduced homology associated with (C, ) and denoted by

Hi#(C) = Hi(C0), i ∈ Z.

Definition 3.2. A chain complex C with augmentation  is acyclic if its corresponding reduced chain complex is acyclic.

Hence the chain complex of an one point space with the above given augmentation is acyclic.

Proposition 3.1. A complex C with augmentation  is acyclic if Hi(C) =

(0 if i 6= 0 R if i = 0.

Proof. Using the fact that Hi#(C) = Hi(C) for i 6= 0, we see that Hi#(C) = 0 if and only if Hi(C) = 0 for i 6= 0.

Since  : C0 → R is surjective, C0/ ker  ∼= R. If C0 is acyclic, ker  = Im ∂1. In this case, C0/ Im ∂1∼= R. Note that H0(C) = C0/ Im ∂1, we find H0(C) ∼= R. This completes the proof.



(8)

4. Category of complexes and the Homological Functor

Let f : X → Y be a continuous map. Given a singular q-simplex σ : ∆q → X in X, the continuous map f ◦ σ : ∆q → Y determine a q-simplex in Y. We define an R-linear map

Cq(f ) : Cq(X, R) → Cq(Y, R), X

σ

rσσ 7→X

σ

rσ(f ◦ σ).

Then we obtain the following diagrams of R-linear maps

· · · −−−−→ Cq+1(X, R)

X

−−−−→ Cq+1 q(X, R)

X

−−−−→ Cq q−1(X, R) −−−−→ · · ·

y Cq+1(f )

y Cq(f )

y Cq−1(f )

 y

 y

· · · −−−−→ Cq+1(Y, R)

Y

−−−−→ Cq+1 q(Y, R)

Y

−−−−→ Cq q−1(Y, R) −−−−→ · · · .

Here ∂qX and ∂qY are the boundaries maps on Cq(X, R) and on Cq(Y, R) respectively. If σ : ∆q→ X is a singular q-simplex, then ∂qσ =Pq

i=0(−1)iσ(i). Thus Cq−1(f ) ◦ ∂qX(σ) = Cq−1(f )(∂qσ) =

q

X

i=0

(−1)if ◦ σ(i)=

q

X

i=0

(−1)i(f ◦ σ)(i). On the other hand, Cq(f )(σ) = f ◦ σ and hence

qYCq(f )(σ) =

q

X

i=0

(−1)i(f ◦ σ)(i). The above two equations imply that

qYCq(f ) = Cq−1(f )∂qX.

In other words, the above diagram commutes. This motivates the definition of chain maps or morphisms between chain complexes.

Definition 4.1. Let A = (Ai, ∂iA) and B = (Bi, ∂iB) be two chain complexes over R. A chain map f = {fi} (morphism) from A to B is a sequence of R-linear maps fi: Ai→ Bi such that

iB◦ fi= fi−1◦ ∂iA, for all i ∈ Z.

In other words, the following diagram commutes:

· · · −−−−→ Ai+1

A

−−−−→ Ai+1 i A

−−−−→ Ai i−1 −−−−→ · · ·

y fi+1

y fi

y fi−1

 y

 y

· · · −−−−→ Bi+1

Bi+1

−−−−→ Bi

Bi

−−−−→ Bi−1 −−−−→ · · · .

The identify morphism idA: A → A is defined by the sequence of identity maps {idAi: Ai→ Ai}.

If f : A → B and g : B → C are chain maps, we define their composition g ◦ f to be the sequence of R-linear maps {gi◦ fi: Ai→ Ci}. We leave to the reader to verify that g ◦ f is again a chain map.

Proposition 4.1. The collections of all complexes over R together with chain maps forms a category denoted by CompR. The category CompR is called the category of complexes over R.

Corollary 4.1. Any continuous map f : X → Y between topological spaces X, Y determines a chain map

C(f ) : C(X, R) → C(Y, R),

between the singular chain complexes of X and Y , where C(f ) = {Ci(f ) : i ∈ Z} such that the following properties hold.

(1) If idX : X → X is the identity map, then C(idX) = idC(X,R).

(2) If f : X → Y and Y → Z are continuous maps, then C(g ◦ f ) = C(g) ◦ C(f ).

(9)

This corollary implies that C: Top → CompR defines a functor from the category of topological spaces into the category of complexes over R.

Lemma 4.1. Let f : A → A0 be a chain map between chain complexes A and B. For each i ∈ Z, we define

Hi(f ) : Hi(A) → Hi(B), [z] 7→ [fi(z)].

Hi(f ) is a well-defined R-linear homomorphism.

Proof. Let z, z0 ∈ ker ∂i so that z − z0 = ∂Ai+1w for some w ∈ Ai+1. Since f is a chain map, fi(z) − fi(z0) = fi(z − z0) = fii+1A w = ∂i+1B fi+1w

Hence fi(z) − fi(z0) ∈ Im ∂i+1B . Thus [fi(z)] = [fi(z0)]. The linearity of Hi(f ) follows from the

linearity of fi. 

Lemma 4.2. Let f : A → B and g : B → C be chain maps. Then Hi(g ◦ f ) = Hi(g) ◦ Hi(f ) for any i ∈ Z.

Proof. Routine check. 

Corollary 4.2. For each i ∈ Z, the assignment Hi : A 7→ Hi(A) sending a chain complex over R to its i-th homology and Hi : (f : A → B) 7→ Hi(f ) : Hi(A) → Hi(B) sending a chain map to its corresponding induced map defines a functor

Hi: CompR→ ModR

from the category of complexes over R to the category of R-modules ModR.

(10)

5. Homotopy and Chain Homotopy

參考文獻

相關文件

(29) Identify the intervals on which the following function are concave up and concave down, decreasing and increasing.. (7) Identify the intervals on which the function are

[This function is named after the electrical engineer Oliver Heaviside (1850–1925) and can be used to describe an electric current that is switched on at time t = 0.] Its graph

We complete the proof of

Hence if we know g, the restriction of f to the boundary of S 1 , and express g as an infinite series of the form (1.2), then we can find (solve for) f.. We need the notion of

In fact, we can drop this

Generalization Theorem Let f be integrable on K = [a, b] × [c, d] to R and suppose that for each y ∈ [c, d], the function x 7→ f (x, y) of [a, b] into R is continuous except

Using the property of Noetherian space, we can decompose any algebraic set in A n (k) uniquely into a finite union of its irreducible components.. Hence we

it so that the corner point touch the upper edge as shown in the figure.. Find x such that the area A