INTRODUCTION TO ALGEBRAIC TOPOLOGY MARTIN ˇCADEK 3. Simplicial and singular homology 3.1. Exact sequences. A sequence of homomorphisms of Abelian groups or modules over a ring . . . fn+1 −−−→ An fn −−→ An−1 fn−1 −−−→ An−2 fn−2 −−−→ . . . is called an exact sequence if Im fn = Ker fn−1. Exactness of the following sequences O −→ A f −−→ B, B g −−→ C −→ 0, 0 −→ C h −−→ D −→ 0 means that f is a monomorphism, g is an epimorphism and h is an isomorphism, respectively. A short exact sequence is an exact sequence 0 −→ A i −→ B j −→ C −→ 0. In this case C ∼= B/A. We say that the short exact sequence splits if one of the following three equivalent conditions is satisfied: (1) There is a homomorphism p : B → A such that pi = idA. (2) There is a homomorphism q : C → B such that jq = idC. (3) There are homomorphisms p : B → A and q : C → B such that ip + qj = idB. The last condition means that B ∼= A ⊕ C with isomorphism (p, q) : B → A ⊕ C. Exercise. Prove the equivalence of (1), (2) and (3). 3.2. Chain complexes. The chain complex (C, ∂) is a sequence of Abelian groups (or modules over a ring) and their homomorphisms indexed by integers . . . ∂n+2 −−−→ Cn+1 ∂n+1 −−−→ Cn ∂n −−→ Cn−1 ∂n−1 −−−→ . . . such that ∂n−1∂n = 0. This conditions means that Im ∂n ⊆ Ker ∂n−1. The homomorphism ∂n is called a boundary operator. A chain homomorphism of chain complexes (C, ∂C ) and (D, ∂D ) is a sequence of homomorphisms of Abelian groups (or modules over a ring) fn : Cn → Dn which commute with the boundary operators ∂D n fn = fn−1∂C n . 1 2 3.3. Homology of chain complexes. The n-th homology group of the chain complex (C, ∂) is the group Hn(C) = Ker ∂n Im ∂n+1 . The elements of Ker ∂n = Zn are called cycles of dimension n and the elements of Im ∂n+1 = Bn are called boundaries (of dimension n). If a chain complex is exact, then its homology groups are trivial. The component fn of the chain homomorphism f : (C, ∂C ) → (D, ∂D ) maps cycles into cycles and boundaries into boundaries. It enables us to define Hn(f) : Hn(C) → Hn(D) by the prescription Hn(f)[c] = [fn(c)] where [c] ∈ Hn(C∗) and [fn(c)] ∈ Hn(D∗ ) are classes represented by the elements c ∈ Zn(C) and fn(c) ∈ Zn(D), respectively. 3.4. Long exact sequence in homology. A sequence of chain homomorphisms . . . −→ A f −−→ B g −−→ C −→ . . . is exact if for every n ∈ Z . . . −→ An fn −−→ Bn gn −−→ Cn −→ . . . is an exact sequence of Abelian groups. Theorem. Let 0 → A i −→ B j −→ C → 0 be a short exact sequence of chain complexes. Then there is a connecting homomorphism ∂∗ : Hn(C) → Hn−1(A) such that the sequence . . . ∂∗ −−→ Hn(A) Hn(i) −−−→ Hn(B) Hn(j) −−−−→ Hn(C) ∂∗ −−→ Hn−1(A) Hn−1(i) −−−−−→ . . . is exact. Proof. Define the connecting homomorphism ∂∗. Let [c] ∈ Hn(C) where c ∈ Cn is a cycle. Since j : Bn → Cn is an epimorphism, there is b ∈ Bn such that j(b) = c. Further, j(∂b) = ∂j(b) = ∂c = 0. From exactness there is a ∈ An−1 such that i(a) = ∂b. Since i(∂a) = ∂i(a) = ∂∂b = 0 and i is a monomorphism, ∂a = 0 and a is a cycle in An−1. Put ∂∗[c] = [a]. Now we have to show that the definition is correct, i. e. independent of the choice of c and b, and to prove exactness. For this purpose it is advantageous to use an appropriate diagram. It is not difficult and we leave it as an exercise to the reader. 3.5. Chain homotopy. Let f, g : C → D be two chain homomorphisms. We say that they are chain homotopic if there are homomorphisms sn : Cn → Dn+1 such that ∂D n+1sn + sn−1∂C n = fn − gn for all n. The relation to be chain homotopic is an equivalence. The sequence of maps sn is called a chain homotopy. 3 Theorem. If two chain homomorphism f, g : C → D are chain homotopic, then Hn(f) = Hn(g). Exercise. Prove the previous theorem from the definitions. 3.6. Five Lemma. Consider the diagram A // f1 ∼=  B // f2 ∼=  C // f3  D // f4 ∼=  E f5 ∼=  ¯A // ¯B // ¯C // ¯D // ¯E If the horizontal sequences are exact and f1, f2, f4 and f5 are isomorphisms, then f3 is also an isomorphism. Exercise. Prove 5-lemma. 3.7. Simplicial homology. We describe two basic ways how to define homology groups for topological spaces – simplicial homology which is closer to geometric intuition and singular homology which is more general. For the definition of simplicial homology we need the notion of ∆-complex, which is a special case of CW-complex. Let v0, v1, . . . , vn be points in Rm such that v1 − v0, v2 − v0, vn − v0 are linearly independent. The n-simplex [v0, v1, . . . , vn] with the vertices v0, v1, . . . , vn is the subspace of Rm { n i=0 tivi; n i=1 ti = 1, ti ≥ 0} with a given ordering of vertices. A face of this simplex is any simplex determined by a proper subset of vertices in the given ordering. Let ∆α, α ∈ J be a collection of simplices. Subdivide all their faces of dimension i into sets Fi β. A ∆-complex is a quotient space of disjoint union α∈J ∆α obtained by identifying simplices from every Fi β into one single simplex via affine maps which preserve the ordering of vertices. Thus every ∆-complex is determined only by combinatorial data. A special case of ∆-complex is a finite simplicial complex. It is a union of simplices the vertices of which lie in a given finite set of points {v0, v1, . . . , vn} in Rm such that v1 − v0, v2 − v0, . . . , vn − v0 are linearly independent. Example. Torus, real projective space of dimension 2 and Klein bottle are ∆-complexes as one can see from the following pictures. In all the cases we have two sets F2 whose elements are triangles, three sets F1 every with two segments and one set F0 containing all six vertices of both triangles. These surfaces are also homeomorhic to finite simplicial complexes, but their structure as simplicial complexes is more complicated than their structure as ∆-complexes. To every ∆-complex X we can assign the chain complex (C, ∂) where Cn(X) is a free Abelian group generated by n-simplices of X (i. e. the rank of Cn(X) is the number 4 b b b b b b a a a a a a c c c Figure 3.1. Torus, RP2 and Klein bottle as ∆-complexes of the sets Fn and the boundary operator on generators is given by ∂[v0, v1, . . . , vn] = n i=0 (−1)i [v0, . . . , ˆvi . . . , vn]. Here the symbol ˆvi means that the vertex vi is omitted. Prove that ∂∂ = 0. The simplicial homology groups of ∆-complex X are the homology groups of the chain complex defined above. Later, we will show that these groups are independent of ∆-complex structure. Exercise. Compute simplicial homology of S2 (find a ∆-complex structure), RP2 , torus and Klein bottle (with ∆-complex structures given in example above). Let X and Y be two ∆-complexes and f : X → Y a map which maps every simplex of X into a simplex of Y and it is affine on all simplexes. Using appropriate sign conventions we can define the chain homomorphism fn : Cn(X) → Cn(Y ) induced by the map f. This chain map enables us to define homomorphism of simplicial homology groups induced by f. Having a ∆-subcomplex A of a ∆-complex X (i. e. subspace of X formed by some of the simplices of X) we can define simplicial homology groups Hn(X, A). The definition is the same as for singular homology in paragraph 3.9. These groups fit into the long exact sequence · · · → Hn(A) → Hn(X) → Hn(X, A) → Hn−1(A) → . . . See again 3.9. 3.8. Singular homology. The standard n-simplex is the n-simplex ∆n = {(t0, t1, . . . , tn) ∈ Rn+1 ; n i=0 ti = 1; ti ≥ 0}. The j-th face of this standard simplex is the (n−1)-dimensional simplex [e0, . . . , ˆej, . . . , en] where ej is the vertex with all coordinates 0 with the exception of the j-th one which is 1. Define εj n : ∆n−1 → ∆n 5 as the affine map εj n(t0, t1, . . . , tn−1) = (t0, . . . , tj−1, 0, tj, . . . , tn−1) which maps e0 → e0, . . . , ej−1 → ej−1, ej → ej+1, . . . , en−1 → en. It is not difficult to prove Lemma. εk n+1εj n = εj+1 n+1εk n for k < j. A singular n-simplex in a space X is a continuous map σ : ∆n → X. Denote the free Abelian group generated by all the singular n-simplices by Cn(X) and define the boundary operator ∂n : Cn(X) → Cn−1(X) by ∂n(σ) = n i=0 (−1)i σεi n for n ≥ 0. Put Cn(X) = 0 for n < 0. Using the lemma above one can show that ∂n+1∂n = 0. The chain complex (Cn, ∂n) is called the singular chain complex of the space X. The singular homology groups Hn(X) of the space X are the homology groups of the chain complex (Cn(X), ∂n), i. e. Hn(X) = Ker ∂n Im ∂n+1 . Next consider a map f : X → Y . Define the chain homomorhism Cn(f) : Cn(X) → Cn(Y ) on singular n-simplices as the composition Cn(f)(σ) = fσ. From definitions it is easy to show that these homomorphisms commute with boundary operators. Hence this chain homomorphism induces homomorphisms f∗ = Hn(f) : Hn(X) → Hn(Y ). Moreover, Hn(idX) = idHn(X) and Hn(fg) = Hn(f)Hn(g). It means that Hn is a functor from the category Top to the category Ab of Abelian groups and their homomorphisms. This functor is the composition of the functor C from Top to chain complexes and the n-th homology functor from chain complexes to abelian groups. Prove the lemma above and ∂n+1∂n = 0. Show directly from the definition that the singular homology groups of a point are H0(∗) = Z and Hn(∗) = 0 for n = 0. 3.9. Singular homology groups of a pair. Consider a pair of topological spaces (X, A). Then the Cn(A) is a subgroup of Cn(X). Hence we get this short exact sequence 0 → Cn(A) i −→ Cn(X) j −→ Cn(X) Cn(A) → 0. Since the boundary operators in Cn(A) are restrictions of boundary operators in Cn(X), we can define boundary operators ∂n : Cn(X) Cn(A) → Cn−1(X) Cn−1(A) . 6 We will denote this chain complex as (C(X, A), ∂) and its homology groups as Hn(X, A). Notice that the factor Cn(X)/Cn(A) is a free Abelian group generated by singular simplices σ : ∆n → X such that σ(∆n ) A. We will need it later. A map f : (X, A) → (Y, B) induces the chain homomorphism Cn(f) : Cn(X) → Cn(Y ) which restricts to a chain homomorphism Cn(A) → Cn(B) since f(A) ⊆ B. Hence we can define the chain homomorphism Cn(f) : Cn(X, A) → Cn(Y, B) which in homology induces the homomorphism f∗ = Hn(f) : Hn(X, A) → Hn(Y, B). We can again conclude that Hn is a functor from the category Top2 into the category Ab of Abelian groups. This functor extends the functor defined on the category Top since every object X and every morphism f : X → Y in Top can be considered as the object (X, ∅) and the morphism ˆf = f : (X, ∅) → (Y, ∅) in the category Top2 and Hn(X, ∅) = Hn(X), Hn( ˆf) = Hn(f). 3.10. Long exact sequence for singular homology. Consider inclusions of spaces i : A → X, i : B → Y and maps j : (X, ∅) → (X, A), j : (Y, ∅) → (Y, B) induced by idX and idY , respectively. Let f : (X, A) → (Y, B) be a map. Then there are connecting homomorphisms ∂X ∗ and ∂Y ∗ such that the following diagram ∂X ∗ // Hn(A) i∗ // (f/A)∗  Hn(X1 ) j∗ // f∗  Hn(X, A) ∂X ∗ // f∗  Hn−1(A) i∗ // (f/A)∗ ∂Y ∗ // Hn(B) i∗ // Hn(Y ) j∗ // Hn(Y, B) ∂Y ∗ // Hn−1(B) i∗ // commutes and its horizontal sequences are exact. An analogous theorem holds also for simplicial homology. Remark. Consider the functor I : Top2 → Top2 which assigns to every pair (X, A) the pair (A, ∅). The commutativity of the last square in the diagram above means that ∂∗ is a natural transformation of functors Hn and Hn−1 ◦ I defined on Top2 . Proof. We have the following commutative diagram of chain complexes 0 // C(A) C(i) // C(f/A)  C(X) C(j) // C(f)  C(X, A) // C(f)  0 0 // C(B) C(i ) // C(Y ) C(j ) // C(Y, B) // 0 with exact horizontal rows. Then Theorem 3.4 and the construction of connecting homomorphism ∂∗ imply the required statement. Remark. It is useful to realize how ∂∗ : Hn(X, A) → Hn−1(A) is defined. Every element of Hn(X, A) is represented by a chain x ∈ Cn(X) with a boundary ∂x ∈ 7 Cn−1(A). This is a cycle in Cn(A) and from the definition in 3.4 we have ∂∗[x] = [∂x]. 3.11. Homotopy invariance. If two maps f, g : (X, A) → (Y, B) are homotopic, then they induce the same homomorphisms f∗ = g∗ : Hn(X, A) → Hn(Y, B). Proof. We need to prove that the homotopy between f and g induces a chain homotopy between C∗(f) and C∗(g). For the proof see [Hatcher], Theorem 2.10 and Proposition 2.19 or [Spanier], Chapter 4, Section 4. Corollary. If X and Y are homotopy equivalent spaces, then Hn(X) ∼= Hn(Y ). 3.12. Excision Theorem. There are two equivalent versions of this theorem. Theorem (Excision Theorem, 1st version). Consider spaces C ⊆ A ⊆ X and suppose that ¯C ⊆ int A. Then the inclusion i : (X − C, A − C) → (X, A) induces the isomorphism i∗ : Hn(X − C, A − C) ∼= −→ Hn(X, A). Theorem (Excision Theorem, 2nd version). Consider two subspaces A and B of a space X. Suppose that X = int A ∪ int B. Then the inclusion i : (B, A ∩ B) → (X, A) induces the isomorphism i∗ : Hn(B, A ∩ B) ∼= −→ Hn(X, A). The second version of Excision Theorem holds also for simplicial homology if we suppose that A and B are ∆-subcomplexes of a ∆-complex X and X = A∪B. In this case the proof is easy since the inclusion Cn(i) : Cn(B, A ∩ B) → Cn(A ∪ B, A) is an isomorphism, namely the both chain complexes are generated by the same n- simplices. Exercise. Show that the theorems above are equivalent. The proof of Excision Theorem for singular homology can be found in [Hatcher], pages 119 – 124, or in [Spanier], Chapter 4, Sections 4 and 6. The main step (a little bit technical for beginners) is to prove the following lemma which we will need later. 8 Lemma. Let U = {Uα; α ∈ J} be a collection of subsets of X such that X = α∈J int Uα. Denote the free chain complex generated by singular simplices σ with σ(∆n ) ∈ Uα for some α as CU n (X). Then CU n (X)) → Cn(X) induces isomorphism in homology. Proof of Excision Theorem. Consider U = {A, B}. Then the inclusion Cn(i) : Cn(B, A ∩ B) → CU n (X) Cn(A) is an isomorphism and, moreover, according to the previous lemma, the homology of the second chain complex is Hn(X, A). 3.13. Homology of disjoint union. Let X = α∈J Xα be a disjoint union. Then Hn(X) = α∈J Hn(Xα). The proof follows from the definition and connectivity of σ(∆n ) in X for every singular n-simplex σ. 3.14. Reduced homology groups. For every space X = ∅ we define the augmented chain complex ( ˜C(X), ˜∂) as follows ˜Cn(X) = Cn(X) for n = −1, Z for n = −1. with ˜∂n = ∂n for n = 0 and ∂0( j i=1 niσi) = j i=1 ni. The reduced homology groups ˜Hn(X) are the homology groups of the augmented chain complex. From the definition it is clear that ˜Hn(X) = Hn(X) for n = 0 and ˜Hn(∗) = 0 for all n. For pairs of spaces we define ˜Hn(X, A) = Hn(X, A) for all n. Then theorems on long exact sequence, homotopy invariance and excision hold for reduced homology groups as well. Considering a space X with distinguished point ∗ and applying the long exact sequence for the pair (X, ∗), we get that for all n ˜Hn(X) = ˜Hn(X, ∗) = Hn(X, ∗). Using this equality and the long exact sequence for unreduced homology we get that H0(X) ∼= H0(X, ∗) ⊕ H0(∗) ∼= ˜H0(X) ⊕ Z. 9 Lemma. Let (X, A) be a pair of CW-complexes, X = ∅. Then ˜Hn(X/A) = Hn(X, A) and we have the long exact sequence · · · → ˜Hn(A) → ˜Hn(X) → ˜Hn(X/A) → ˜Hn−1(A) → . . . Proof. According to example in Section 2 (X, A) → (X ∪ CA, CA) → (X ∪ CA/CA, ∗) = (X/A, ∗) is the composition of an excision and a homotopy equivalence. Hence ˜Hn(X/A) = Hn(X, A). The rest folows from the long exact sequence of the pair (X, A). Exercise. Prove that ˜Hn( Xα) ∼= ⊕ ˜Hn(Xα). ˜Hn can be considered as a functor from Top∗ to Abelian groups. 3.15. The long exact sequence of a triple. Three spaces (X, B, A) with the property A ⊆ B ⊆ X are called a triple. Denote i : (B, A) → (X, A) and j : (X, A) → (X, B) maps induced by the inclusion B → X and idX, respectively. Analogously as for pairs one can derive the following long exact sequence: . . . ∂∗ −→ Hn(B, A) i∗ −→ Hn(X, A) j∗ −→ Hn(X, B) ∂∗ −→ Hn−1(B, A) i∗ −→ . . . 3.16. Singular homology groups of spheres. Consider the long exact sequence of the triple (∆n , ∂∆n , Λn−1 = ∂∆n − ∆n−1 ): · · · → Hi(∆n , Λn−1 ) → Hi(∆n , ∂∆n ) ∂∗ −→ Hi−1(∂∆n , Λn−1 ) → Hi−1(∆n , Λn−1 ) → . . . The pair (∆n , Λn−1 ) is homotopy equivalent to (∗, ∗) and hence its homology groups are zeroes. Next using Excision Theorem and homotopy invariance we get that Hi(∆n , Λn−1 ) ∼= Hi(∆n−1 , ∂∆n−1 ). Consequently, we get an isomorphism Hi(∆n , ∂∆n ) ∼= Hi−1(∆n−1 , ∂∆n−1 ). Using induction and computing Hi(∆1 , ∂∆1 ) = Hi([0, 1], {0, 1}) ∼= Hi−1({0, 1}, {0}) we get that Hi(∆n , ∂∆n ) = Z for i = n, 0 for i = n. Doing the induction carefully we can find that the generator of the group Hn(∆n , ∂∆n ) = Z is determined by the singular n-simplex id∆n . The pair (Dn , Sn−1 ) is homeomorphic to (∆n , ∂∆n ). Hence it has the same homology groups. Using the long exact sequence for this pair we obtain ˜Hi−1(Sn−1) = Hi(Dn , Sn−1 ) = 0 for i = n, Z for i = n. 10 3.17. Mayer-Vietoris exact sequence. Denote inclusions A∩B → A, A∩B → B, A → X, B → X by iA, iB, jA, jB, respectively. Let C → A ∩ B and suppose that X = int A ∪ int B. Then the following sequence . . . ∂∗ −−→ Hn(A ∩ B, C) (iA∗,iB∗) −−−−−→ Hn(A, C) ⊕ Hn(B, C) jA∗−jB∗ −−−−−→ Hn(X, C) ∂∗ −−→ Hn−1(A ∩ B, C) −→ . . . is exact. Proof. The covering U = {A, B} satisfies conditions of Lemma 3.12. The sequence of chain complexes 0 −→ C(A ∩ B) C(C) i −→ C(A) C(C) ⊕ C(B) C(c) j −→ CU (X) C(C) −→ 0 where i(x) = (x, x) and j(x, y) = x − y is exact. Consequently, it induces a long exact sequence. Using Lemma 3.12 we get that Hn(CU (X), C(C)) = Hn(X, C), which completes the proof. 3.18. Equality of simplicial and singular homology. Let (X, A) be a pair of ∆-complexes. Then the natural inclusion of simplicial and singular chain complexes C∆ (X, A) → C(X, A) induces the isomorphism of simplicial and singular homology groups H∆ n (X, A) ∼= Hn(X, A). Outline of the proof. Consider the long exact sequences for the pair (Xk , Xk−1 ) of skeletons of X. We get H∆ n+1(Xk , Xk−1 ) //  H∆ n (Xk−1 ) //  H∆ n (Xk ) //  H∆ n (Xk , Xk−1 ) //  H∆ n−1(Xk−1 )  Hn+1(Xk , Xk−1 ) // Hn(Xk−1 ) // Hn(Xk ) // Hn(Xk , Xk−1 ) // Hn−1(Xk−1 ) Using induction by k we have H∆ i (Xk−1 ) = Hi(Xk−1 ) for all i. Further, C∆ i (Xk , Xk−1 ) is according to definition zero if i = k and free Abelian of rank equal the number of isimplices ∆i α if i = k. The homology groups H∆ i (Xk , Xk−1 ) have the same description. Since α ∆k α/ α ∂∆k α = Xk /Xk−1 we get the isomorphism H∆ i (Xk /Xk−1 ) → Hi( α ∆k α/ α ∂∆k α) = Hi(Xk /Xk−1 ). Applying 5-lemma (see 3.6) in the diagram above, we get that H∆ n (Xk ) → Hn(Xk ) is an isomorphism. If X is finite ∆-complex, we are ready. If it is not, we have to prove that H∆ n (X) = Hn(X). See [Hatcher], page 130. 11 CZ.1.07/2.2.00/28.0041 Centrum interaktivních a multimediálních studijních opor pro inovaci výuky a efektivní učení