用户: Eee/Hatcher

For a quick review of Homology Theory, one can refer to the Appendix of Characteristic Classes written by Milnor and Stasheff.

0Some Underlying Geometric Notions

Definition 0.1. A retraction of onto a subspace is a map such that .
A deformation retraction of onto is a homotopy from to .

Definition 0.2. For a subspace and a map , , with attached along via , is the quotient space .

Definition 0.3 (Operations on Spaces).
The cone .
The suspension .
The mapping cylinder is .
The mapping cone is .
The join , that is, collapsing to and to .
The wedge sum for chosen points .
The smash product .

Definition 0.4. A cell complex or a CW complex is a space obtained by , starting from a set of points and inductively forming the -skeleton by attaching -cells . Either set stopping at some finite , or set giving the weak topology: is open iff is open for each .

Definition 0.5. A subcomplex of is a closed subspace that is a union of cells of . is called a CW pair.

Example 0.6. A graph is , vertices with edges attached, two ends of which can be attached to the same vertex.

Example 0.7. The sphere has the structure .

Example 0.8. Real projective -space is also the quotient space of a semisphere with antipodal points identified (just ), or . So .

Example 0.9. Complex projective -space .

Definition 0.10. Orientable surface

Attach -cell to the wedge sum of circles by the word we obtain an orientable surface .

Definition 0.11. Nonorientable surface

Attach -cell to the wedge sum of circles by the word we obtain a nonorientable surface .

For example, is the projective plane , and is the Klein bottle.

1Fundamental Group

One can often show that two spaces are not homeomorphic by showing that their fundamental groups are not isomorphic, since it will be easy consequence of the definition of the fundamental group that homeomorphic spaces have isomorphic fundamental groups.

Basic Constructions

Definition 1.1. A homotopy of paths in is a family , such that

  1. The endpoints and are independent of .

  2. The associated map defined by is continuous.

When two paths and are connected in this way by a homotopy , they are said to be homotopic. The notation for this is .

Definition 1.2. The relation of homotopy on paths with fixed endpoints in any space is an equivalence relation.

The equivalence class of a path under the equivalence relation of homotopy will be denoted and called the homotopy class of .

Definition 1.3. The set of all homotopy classes of loops at the basepoint is denoted .

is a group with respect to the product . This group is called the fundamental group of at the basepoint .

Definition 1.4. Define a reparametrization of a path to be a composition where is any continuous map such that and .

Reparametrizing a path preserves its homotopy class since via the homotopy , where .

Definition 1.5. Say is a path from to , then we define a change-of-basepoint map by .

Proposition 1.6. The map is an isomorphism.

Thus if is path-connected, the group is, up to isomorphism, independent of the choice of basepoint . In this case the notation is often abbreviated to .

Definition 1.7. A space is called simply-connected if it is path-connected and has trivial fundamental group.

Proposition 1.8. A space is simply-connected iff there is a unique homotopy class of paths connecting any two points in .

Definition 1.9. Given a space , a covering space of consists of a space and a map satisfying the following condition:

  For each point , there is an open neighborhood of in such that is a union of disjoint open sets each of which is mapped homeomorphically onto by .

Such a will be called evenly covered.

There are three easy lifting properties.

Proposition 1.10. For each path starting at a point and each there is a unique lift starting at .

Proposition 1.11. For each homotopy of paths starting at and each there is a unique lifted homotopy of paths starting at .

Proposition 1.12. Given a map and a map lifting , then there is a unique map lifting and restricting to the given on .

Theorem 1.13. is an infinite cyclic group generated by the homotopy class of the loop based at .

In other words, .

Theorem 1.14. Algebraic fundamental theorem

Every nonconstant polynomial with coefficients in has a root in .

Theorem 1.15. Brouwer fixed point theorem in dimension

Every continunous map has a fixed point, that is, a point with .

Remark 1.16. Brouwer fixed point theorem in dimension

Every continunous map has a fixed point, that is, a point with .

Theorem 1.17. Borsuk-Ulam theorem in dimension

For every continuous map there exists a pair of antipodal points and in with .

Remark 1.18. Borsuk-Ulam theorem in dimension

For every continuous map there exists a pair of antipodal points and in with .

Corollary 1.19. There is no one-to-one continuous map from to , so is not homeomorphic to a subspace of .

Corollary 1.20. Whenever is expressed as the union of three closed sets and , then at least one of these sets must contain a pair of antipodal points .

Projecting the four faces of the tetrahedron radially onto the sphere, we obtain a cover of by four closed sets, none of which contains a pair of antipodal points.

Remark 1.21. Assuming the higher-dimensional version of the Borsuk-Ulam theorem, the same arguments show that cannot be covered by closed sets without antipodal pairs of points, though it can be covered by such sets, as the higher-dimensional analog of a tetrahedron shows.

Proposition 1.22. if and are path-connected.

Example 1.23. The Torus: .

Definition 1.24. Suppose is a map taking the basepoint to the basepoint . For brevity we write in this situation.

Then induces a homomorphism defined by composing loops based at with , that is, .

Proposition 1.25. Two basic properties of induced homomorphisms are:

  1. for a composition .

  2. , which is a concise way of saying that the identity map induces the identity map .

These two properties of induced homomorphisms are what makes the fundamental group a functor.

Lemma 1.26. The decomposition of loops subordinated to an open covering.

If a space is the union of a collection of path-connected open sets each containing the basepoint and if each intersection is path-connected, then every loop in at is homotopic to a product of loops each of which is contained in a single .

Theorem 1.27. if .

Corollary 1.28. is not homeomorphic to for .

Remark 1.29. The more general statement that is not homeomorphic to if can be proved in the same way using either the higher homotopy groups or homology groups.

Proposition 1.30. If a space retracts onto a subspace , then the homomorphism induced by the inclusion is injective. If is a deformation retract of , then is an isomorphism.

Definition 1.31. A basepoint-preserving homotopy is the case that for all .

Similarly, we can define homotopy equivalence for spaces with basepoint. One says if there are maps with homotopies and through maps fixing the basepoints.

Proposition 1.32. If is a basepoint-preserving homotopy, then .

Proposition 1.33. If , then .

For homotopy equivalences one does not have to be quite so careful, as the conditions on basepoints can actually be dropped.

Proposition 1.34. If is a homotopy equivalence, then the induced homomorphism is an isomorphism for all .

Lemma 1.35. If is a homotopy and is the path formed by the images of a basepoint , then we have .

Van Kampen’s Theorem

The Van Kampen theorem gives a method for computing the fundamental groups of spaces that can be decomposed into simpler spaces whose fundamental groups are already known.

We shall see for example that for every group there is a space whose fundamental group is isomorphic to .

Definition 1.36. Free product. Refer to Hatcher P41.

Definition 1.37. Free group is the free product of any number of copies of , finite or infinite. The generators are called a basis for the free group, and the number of basis element is the rank of the free group.

Definition 1.38. A basic property of the free product is that any collection of homomorphisms extends uniquely to a homomorphism .

Namely, the value of on a word with must be , and using this formula to define gives a well-defined homomorphism.

Suppose a space is decomposed as the union of a collection of path-connected open subsets , each of which contains the basepoint . The homomorphisms induced by the inclusions extend to a homomorphism .

The Van Kampen theorem will say that is very often surjective, but we can expect to have a nontrivial kernel in general.

For is the homomorphism induced by the inclusion then , both these compositions being induced by the inclusion , so the kernel of contains all the elements of the form for .

Theorem 1.39. Van Kampen theorem

If is a union of path-connected open sets each containing the basepoint and if each intersection is path-connected, then the homomorphism is surjective.

If in addition each intersection is path-connected, then the kernel of is the normal subgroup generated by all elements of the form for , and hence induces an isomorphism .

Example 1.40. Wedge sum: .

Remark 1.41. Van Kampen’s theorem is often applied when there are just two sets and in the cover of , so the condition on the triple intersections is superfluous and one obtains an isomorphism , under the assumption that is path-connected.

Example 1.42. Linking of Circles

The complement of a single circle deformation retracts onto a wedge sum embedded in . Hence .

In similar fashion, the complement of two unlinked circles and deformation retracts onto , hence, .

On the other hand, if and are linked, then deformation retracts onto the wedge sum of and a torus separating and , hence .

For details and figures, please refer to Hatcher P46.

For the remainder of this section we shall be interested in cell complexes, and in particular in how the fundamental group is affected by attaching -cells.

Suppose we attach a collection of -cells to a path-connected space via maps , producing a space . If is a basepoint of then determines a loop at that we shall call .

For different ’s the basepoint of these loops may not all coincide. To remedy this, choose a basepoint and a path in from to for each . Then is a loop at .

This loop may not be nullhomotopic in , but it will certainly be nullhomotopic after the cell is attached. Thus the normal subgroup generated by all the loops for varying lies in the kernel of the map induced by the inclusion .

Proposition 1.43. If is obtained from by attaching -cells as described above, then the inclusion induces a surjection whose kernel is . Thus .

Proposition 1.44. If is obtained from by attaching -cells for a fixed , then the inclusion induces an isomorphism .

For a path-connected cell complex the inclusion of the -skeleton induces an isomorphism .

Remark 1.45. is independent of the choice of the paths . If we replace by another path having the same endpoints, then changes to , so and define conjugate elements of .

Definition 1.46. denotes the group with generators and relator , in other words, the free group on the generator modulo the normal subgroup generated by the words in these generators.

Proposition 1.47.

Corollary 1.48. The surface is not homeomorphic, or even homotopy equivalent, to if .

Proposition 1.49. This abelianizes to the direct sum of with copies of since in the abelianzation we can rechoose the generators to be and with .

Corollary 1.50. is not homotopy equivalent to if , nor is homotopy equivalent to any orientable surface .

Corollary 1.51. For every group , there is a -dimensional cell complex with .

Remark 1.52. However, the cell complex may not be a surface. For more details, please refer to Hatcher P52.

Covering Spaces

The connection between the fundamental group and covering spaces runs much deeper than this, and in many ways they can be regarded as two viewpoints toward the same thing. Algebraic aspects of the fundamental group can often be translated into geometric language of covering spaces.

Definition 1.53. A covering space of a space is a space together with a map satisfying the following condition: Each point has an open neighborhood in such that is a union of disjoint open sets in , each of which is mapped homeomorphically onto by .

Such a is called evenly covered and the disjoint open sets in that project homeomorphically to by are called sheets of over .

We allow to be empty, so need not be surjective. Remark that this definition varies from person to person.

The number of sheets over is the cardinality of for . As varies over this number is locally constant, so it is constant if is connected.

Definition 1.54. A lift of a map is a map such that .

Proposition 1.55. Homotopy lifting property

Given a covering space , a homotopy , and a map lifting , then there exists a unique homotopy of that lifts .

Remark 1.56. Taking to be a point gives the path lifting property.

Proposition 1.57. The map induced by a covering space is injective.

The image subgroup in consists of the homotopy classes of loops in based at whose lift to starting at are loops.

Proposition 1.58. The number of sheets of a covering space with and path-connected equals the index of in .

Proposition 1.59. Lifting criterion

Suppose given a covering space and a map with path-connected and locally path-connected.

Then a lift of exists iff .

Proposition 1.60. Unique lifting property

Given a covering sapce and a map , if two lifts of agree at one point of and is connected, then and agree on all of .

The Classification of Covering spaces

The main thrust of the classification will be a correspondence between connected covering spaces of and subgroups of . This is often called the Galois correspondence because of its surprising similarity to another basic correspondence in the purely algebraic subject of Galois theory.

The Galois correspondence arises from the function that assigns to each covering space the subgroup of . Fisrt we consider whether this function is surjective. That is, we ask whether every subgroup of is realized as for some covering space . In particular we can ask whether the trivial subgroup is realized. Since is always injective, this amounts to asking whether has a simply-connected covering space.

Definition 1.61. Semilocally simply-connected

We say is semilocally simply-connected if each point has a neighborhood such that the inclusion-induced map is trivial.

A locally simply-connected space is certainly semilocally simply-connected. An example of a space that is not semilocally simply-connected is the shrinking wedge of circles.

Given a path-connected, locally path-connected, semilocally simply-connected space with a basepoint , we are therefore led to define Let be the collection of path-connected open sets such that is trivial. Say that is a basis for the topology on if is locally path-connected, semilocally simply-connected.

Given a set and a path in from to a point in , let depends only on the homotopy class and if . And the sets form a basis for a topology basis for a topology on .

The bijection is a homeomorphism. This implies is continuous and then we can also deduce that this is a covering space.

It remains only to show that is simply-connected. Then we complete the construction of a simply-connected covering space .

The hypotheses for constructing a simply-connected covering space of in fact suffice for constructing covering spaces realizing arbitrary subgroup of .

Proposition 1.62. Suppose is path-connected, locally path-connected, semilocally simply-connected. Then for every subgroup there is a covering space such that for a suitably chosen basepoint .

Definition 1.63. An isomorphism between covering spaces and is a homeomorphism such that .

Proposition 1.64. If is a path-connected and locally path-connected, then two path-connected covering spaces and are isomorphic via an isomorphism taking a basepoint to a basepoint iff .

Theorem 1.65. Classification Theorem

Let be a path-connected, locally path-connected, semilocally simply-connected space. Then there is a bijection between the set of basepoint-preserving isomorphism classes of path-connected covering spaces and the set of subgroups of , obtained by associating the subgroup to the covering space .

If basepoints are ignored, this correspondence gives a bijection between isomorphism classes of path-connected covering sapce and conjugacy classes of subgroups of .

Definition 1.66. A simply-connected covering space of a path-connected, locally path-connected space is a covering sapce of every other path-connected covering space of .

A simply-connected covering space of is therefore called a universal cover. It is unique up to isomorphism, so one is justified in calling it the universal covering.

Definition 1.67. For a covering space the isomorphisms are called deck transformations or covering transformations. These form a group under composition.

Example 1.68. projecting a vertical helix onto a circle, in this case.

Example 1.69. The -sheeted covering space , the deck transformations are the rotations of through angles that the multiples of , so .

By the unique lifting property, a deck transformation is completely determined by where it sends a single point, assuming is path-connected. In particular, only the identity deck transformation can fix a point of .

Definition 1.70. A covering space is called normal if for each and each pair of lifts of there is a deck transformation taking to .

For example, the covering space and the -sheeted covering spaces are normal.

Proposition 1.71. Let be a path-connected covering space of the path-connected, locally path-connected space , ane let be the subgroup . Then:

  (a) This covering space is normal iff is a normal subgroup of .

  (b) is isomorphic to the quotient where is the normalizer of in .

In particular, is isomorphic to if is a normal covering. Hence for the universal cover we have .

Definition 1.72. We call an action satisfying the following condition is a covering space action or properly discontinuous action.

Each has a neighborhood such that all the images for varying are disjoint. In other words, implies .

Equivalently, only when is the identity.

Proposition 1.73. The action of the deck transformation group on is a covering space action.

Definition 1.74. The orbits in , and is called the orbit space of the action.

For example, for a normal covering space , the orbit space is just .

Proposition 1.75. If an action of a group on a space satisfies covering space action condition, then

(a) The quotient map , , is a normal covering space.

(b) is the group of deck transformation of this covering space if is path-connected.

(c) is isomorphic to if is path-connected and locally path-connected.

Corollary 1.76. For a covering space action of a group on a simply-connected locally path-connected space , the orbit space has fundamental group isomorphic to .

Example 1.77. A covering sapce . For details please refer to Hatcher P73.

Example 1.78. .

The antipodal map of , , generates an action of on with orbit space , real projective -space.

Since is simply-connected if , we deduce that for .

2Homology

Simplicial and Singular Homology

Simplicial homology

Example 2.1. with one vertex and one edge . Then and are both and the boundary map is zero since . Hence

Example 2.2. , the torus with the -complex structure (a square identifying the opposite edges and we draw one diagonal denoted by .) Denote the -simplices and .

so . Since and is a basis for , it follows that with basis the homology classes and .

Since there are no -complices, is equal to , which is infinite cyclic generated by . Thus

Example 2.3. , the projective plane with the -complex structure (a square identifying the opposite edges with orientation revered and we draw one diagonal denoted by .) Denote the -simplices and , the -simplices and .

The image of is generated by , so with either vertex as a generator.

Since and , we see that is injective, so .

Further, with basis and and is an index-two subgroup of since we can choose and as a basis for and and as a basis of . Thus .

Example 2.4. We can obtain a -complex structure on by taking two copies of and identifying their boundaries via the identity map. Labeling these two -simplices and , then it is obvious that is infinite cyclic generated by . Thus for this -complex structure on .

Singular homology

Definition 2.5. A singular -simplex in a space is by definition just a continuous map .

Remark 2.6. The word “singular” is used here to express the idea that need not be a nice embedding but can have “singularities” where its image does not look at all like a simplex.

Definition 2.7. Let be the free abelian group with basis the set of singular -simplices in . Elements of , called -chains, are finite formal sums for and .

Definition 2.8. Singular homology group .

Remark 2.9. It is evident from the definition that homeomorphic spaces have isomorphic singular homology groups , in contrast with the situation for .

Proposition 2.10. Corresponding to the decomposition of a space into its path-components there is an isomorphism of with the direct sum .

Proposition 2.11. If is nonempty and path-connected, then . Hence for any space , is a direct sum of ’s, one for each path-component of .

Proposition 2.12. If is a point, then for and .

Definition 2.13. Define the reduced homology groups to be the homology groups of the augmented chain complex where .

Remark 2.14. Formally, one can think of the extra in the augmented chain complex as generated by the unique map where is the empty simplex, with no vertices. The augmentation map is then the usual boundary map since .

Proposition 2.15. A chain map between chain complexes induces homomorphisms between the homology groups of the two complexes.

Theorem 2.16. If two maps are homotopic, then they induce the same homomorphism .

Corollary 2.17. The maps induced by a homotopy equivalence are isomorphisms for all .

Proposition 2.18. Chain-homotopic chain maps induce the same homomorphism on homology.

The properties of induced homomorphisms we stated above hold equally well in the setting of reduced homology, with the same proofs.

Exact Sequences and Excision

If there was always a simple relationship between the homology groups of a space , a subspace , and the quotient space , then this could be a very useful tool in understanding the homology groups of spaces such as CW complexes that can be built inductively from successively more complicated subspaces.

The novel feature of the actual relationship is that it involves the groups , and for all values of simultaneously.

In practice this is not as bad as it might sound, and in addition it has the pleasant side effect of sometimes allowing high-dimensional homology groups to be computed in terms of lower-dimensional groups which may already be known, for example by induction.

Theorem 2.19. If is a space and is a nonempty closed subspace that is a deformation retract of some neighborhood in , then there is an exact sequence where is the inclusion and is the quotient map .

Remark 2.20. The map will be constructed in the course of the proof. The idea is that an element can be represented by a chain in with a cycle in whose homology class is .

Definition 2.21. Pairs of spaces satisfying the hypothesis of the theorem will be called good pairs.

For example, if is a CW complex and is a nonempty subcomplex, then is a good pair.

Corollary 2.22. and for .

Proof.
Proof. For , take so . Then we could show that .

Corollary 2.23. Brouwer theorem

is not a retract of . Hence every map has a fixed point.

Definition 2.24. Let be the quotient group . The boundary map is induced by the in .

holds then becomes a complex, the homology group corresponding to this complex is called relative homology groups.

Elements in are called relative cycles, -chains with .

Elements in are called relative boundaries, for and .

Proposition 2.25. The commutative diagram with the columns are exact and the rows are chain complexes which we denote , and . Such a diagram is called a short exact sequence of chain complexes. This short exact sequence of chain complexes stretches out into a long exact sequence of homology groups where denotes the homology group at in the chain complex , and and are defined similarly.

Now define the boundary map . Let be a cycle, and for some as is onto. Since implies , we have for some , where since and is injective. Take .

The theorem represents the beginnings of the subject of homological algebra. The method of proof is sometimes called diagram chasing.

Proposition 2.26. From the last proposition and the diagram we have the following exact sequence

Proposition 2.27. There is a completely analogous long exact sequence of reduced homology groups for a pair with .

In particular, is the same as for all , where .

Example 2.28.

Example 2.29. for all .

Proposition 2.30. If two maps are homotopic through maps of pairs , then .

Proposition 2.31. An easy generalization of the long exact sequence of a pair is the long exact sequence of a triple , where :

Theorem 2.32. Excision Theorem

Given subspaces such that , then the inclusion induces isomorphisms for all .

Equivalently, for subspaces whose interiors cover , the inclusion induces isomorphism for all .

Remark 2.33. In a metric space “smallness” can be defined in terms of diameters, but for general spaces it will be defined in terms of covers.

Definition 2.34. Let be the subgroup of consisting of chains such that each has image contained in some set in the cover .

Proposition 2.35. The inclusion is a chain homotopy equivalence, that is, there is a chain map such that and are chain homotopic to the identity.

Hence induces isomorphisms for all .

Proposition 2.36. For good pairs , the quotient map induces isomorphisms for all .

Remark 2.37. This proposition shows that relative homology can be expressed as reduced absolute homology in the case of good pairs , but in fact there is a way of doing this for arbitrary pairs: where is the tip of the cone.

Example 2.38. The identity map , viewed as a singular -simplex, is a cycle generating . We claim there are isomorphisms where is the union of all but one of the -dimensional faces of .

Example 2.39. To find a cycle generating let us regard as two -simplices and with their boundaries identified in the obvious way, preserving the ordering of vertices.

The difference , viewed as a singular -chain, is then a cycle, and we claim it represents a generator of . To see this, consider the isomorphisms

Corollary 2.40. If the CW complex is the union of subcomplexes and , then the inclusion induces isomorphisms for all .

Corollary 2.41. For a wedge sum , the inclusions induce an isomorphism , provided that the wedge sum is formed at basepoints such that the pairs are good.

Proof.
Proof. Since reduced homology is the same as homology relative to a basepoint, this follows from the proposition by taking .

Theorem 2.42. If nonempty open sets and are homeomorphic, then .

Proof.
Proof. For we have by excision, the long exact sequence for , and deformation retraction respectively.

Definition 2.43. The local homology groups of a space at a point are defined to be the groups .

For any open neighborhood of , excision gives isomorphisms assuming points are closed in , and thus the groups depend only on the local topology of near .

A homeomorphism must induce isomorphisms for all and , so the local homology groups can be used to tell when spaces are not locally homeomorphic at certain points.

Proposition 2.44. Naturality

For six chain complexes with for any . Moreover, there are three chain maps , , . Then the induced diagram is commutative.

Proposition 2.45. For , the diagram is commutative.

Proposition 2.46. Recall Theorem 2.19, if are spaces and are nonempty closed subspaces that are deformation retracts of some neighborhoods in respectively, then the following diagram commutes

The Equivalence of Simplicial and Singular Homology

Theorem 2.47. The homomorphisms are isomorphisms for all and all -complex pairs .

Lemma 2.48. The Five-Lemma

In a commutative diagram of abelian groups as shown, if the two rows are exact and and are isomorphisms, then is an isomorphism also.

Corollary 2.49. We can deduce from this theorem that is finitely generated whenever is a -complex with finitely many -simplices, since in this case the simplicial chain group is finitely generated, hence also its subgroup of cycles and therefore also the latter group’s quotient .

Definition 2.50. If we write as the direct sum of cyclic groups, then the number of summands is known traditionally as the Betti number of , and integers specifying the orders of the finite cyclic summands are called torsion coefficients.

Degree

Definition 2.51. For a map with , the induced map is a homomorphism from an infinite cyclic group to itself and so must be of the form for some integer depending only on . This integer is called the degree of , with the notation .

Proposition 2.52. Here are some basic properties of degree:

(a)

.

(b)

if is not surjective. (Hint: .)

(c)

If then since . The converse statement, that if , is a fundamental theorem of Hopf from around 1925.

(d)

since . As a consequence, if is a homotopy equivalence.

(e)

if is a reflection of .

(f)

where is the antipodal map which is the composition of reflections.

(g)

If has no fixed points, then since it is not difficult to prove .

Theorem 2.53. has a continuous field of nonzero tangent vectors iff is odd.

Proof.
Proof. Suppose that the tangent vectors , orthogonal to , are all of length . The homotopy between and implies .

Proposition 2.54. is the only nontrivial group that can act freely on if is even.

Proof.
Proof. Since homeomorphisms have degree , an action of a group on determines a degree function . This is a homomorphism since . If the action is free, sends each nontrivial element of to by property (g) above. Thus, when is even, has trivial kernel, so .

Cellular Homology

Lemma 2.55. If is a CW complex, then:

(a)

is zero for and is free abelian for , with a basis in one-to-one correspondence with the -cells of .

(b)

for . In particular, if is finite-dimensional then for .

(c)

The map induced by the inclusion is an isomorphism for and surjective for .

Proof.

Proof.

(a)

.

(b,c)

Consider If , the last term is so is surjective; if , both the first term and last term are so .
Therefore for , for , and is surjective for .

Proposition 2.56. Let be a CW compelx. Using 2.55, portions of the long exact sequence for the pairs , and fit into a diagram where and are defined as the compositions and , which are just ‘relativizations’ of the boundary maps and .

Definition 2.57. Note , the horizontal row in the diagram is a chain complex, called the cellular chain complex of . Since is free with basis in one-to-one correspondence with the -cells of , one can think of elements of as linear combinations of -cells of .

The homology groups of this cellular chain complex are called the cellular homology groups of , temporarily denoted by .

Theorem 2.58. .

Proposition 2.59. Here are a few immediate applications:

(i)

if is a CW complex with no -cells.

(ii)

More generally, if is a CW complex with -cells, then is generated by at most elements.

(iii)

If is a CW complex having no two of its cells in adjacent dimensions, then is free abelian with basis in one-to-one correspondence with the -cells of .

Proof.

Proof.

(ii)

and .

(iii)

The maps are automatically zero.

Example 2.60.

Example 2.61. with , using the product structure consisting of a -cell, two -cells, and a -cell.

Proposition 2.62. Cellular Boundary Formula

where is the degree of the map that is the composition of the attaching map of with the quotient map collapsing to a point.

Remark 2.63. In case is connected and has only one -cell, then must be , otherwise would not be .

Example 2.64. : one -cell, -cells, and one -cell attached by the product of commutators . The associated cellular chain complex is Note that both and are , then one can get

Example 2.65. : one -cell, -cells, and one -cell attached by the word . The associated cellular chain complex is Note that and , then one can get

Remark 2.66. These two examples illustrate the general fact that the orientablility of a close connected manifold of dimension is detected by , which is if is oriented and otherwise. This will be shown in the next chapter.

Definition 2.67. If satisfies for all , we call an acyclic space.

Example 2.68. The examples of and are also amazing. One can refer to Hatcher P142.

Definition 2.69. Moore Spaces

Given an abelian group and an integer , we can construct a CW complex such that and for , commonly written to indicate the dependence of and .

Example 2.70. Real Projective Space

, thus the cellular chain complex for is From this it follows that

Euler Characteristic

Definition 2.71. For a finite CW complex , the Euler characteristic is defined to be the alternating sum where is the number of -cells of .

Remark 2.72. The following result shows that can be defined purely in terms of homology, and hence depends only on the homotopy type of . In particular, is independent of the choice of CW structure on .

Theorem 2.73. .

Definition 2.74. Here the rank of a finitely generated abelian group is the number of summands when the group is expressed as a direct sum of cyclic groups.

If is a short exact sequence of finitely generated abelian groups, then .

Example 2.75. and .

Split Exact Sequences

Lemma 2.76. Splitting Lemma

  For a short exact sequence of abelian groups the following statements are equivalent:

(a)

There is a homomorphism such that .

(b)

There is a homomorphism such that .

(c)

There is a an isomorphism making a commutative diagram, where the maps in the lower row are the obvious ones, and .

  If these conditions are satisfied, the exact sequence is said to be split.

Proposition 2.77. A retraction gives a splitting .

Example 2.78. The last proposition can be used to show the nonexistence of such a retraction in some cases.

(1)

For the Brouwer fixed point theorem, a retraction would give

(2)

The mapping cylinder of a degree map with . If retracted onto the corresponding to the domain of , we would have a split short exact sequence But this sequence does not split since is not isomorphic to if , so the retraction cannot exist.

  The simplest case of degree map , , this says that the Möbius band does not retract onto its boundary circle.

Mayer-Vietoris Sequences

Proposition 2.79. For a pair of subspaces such that is the union of the interiors of and , this exact sequence has the form where .

Proposition 2.80. There is also a formally identical Mayer-Vietoris sequence for reduced homology groups, obtained by augmenting the previous short exact sequence of chain complexes in the obvious way:

Remark 2.81. Mayer-Vietoris sequence can be viewed as analogs of the van Kampen theorem since if is path-connected, the terms of the reduced Mayer-Vietoris sequence yield an isomorphism . This is exactly the abelianized statement of the van Kampen theorem, and is the abelianization of for path-connected spaces.

Remark 2.82. There are also Mayer-Vietoris sequences for decompositions such that and are deformation retracts of neighborhoods and with deformation retracting onto .

Example 2.83. Take with and the northern and southern hemispheres, so that . Then in the reduced Mayer-Vietoris sequence the terms are zero, so we obtain isomorphisms .

Example 2.84. We can decompose the Klein bottle as the union of two Möbius bands and glued together by a homeomorphism between their boundary circles.

Then and are homotopy equivalent to circles, so the interesting part of the reduced Mayer-Vietoris sequence for the decomposition is the segment The map is , , since the boundary circle of a Möbius band wraps twice around the core circle. Since is injective we obtain . Furthermore, we have since .

Homology with coefficients

Definition 2.85. The generalization consists of using chains of the form where each is a singular -simplex in as before, but now the coefficients are taken to be in a fixed abelian group rather than . Such -chains form an abelian group and there is the expected relative version .

   defined similarly satisfies and hence and form chain complexes. The resulting homology groups and are called homology groups with coefficients in .

  Reduced groups are defined via the augmented chain complex with again defined by summing coefficients.

Proposition 2.86. All the theory we developed for coefficients carries over directly to general coefficient group .

Lemma 2.87. If has degree , then is multiplication by .

  Then one could define in the cellular homology.

Proof.
Proof.

Remark 2.88. The case is particularly simple since one is just considering sums of singular simplices with coefficients or , so by discarding terms with coefficient one can think of chains as just finite ‘unions’ of singular simplices.

  The boundary formulas also simplify since one on longer has to worry about signs. Since signs are an algebraic representation of orientation considerations, one can also ignore orientations. This means that homology with coefficients is often the most natural tool in the absence of orientability.

Example 2.89. It is instructive to see what happens to the homology of when the coefficient group is chosen to be a field . The cellular chain complex is Hence if has characteristic , for example , then for , a more uniform answer than with coefficients.

  On the other hand, if has characteristic different from , then the boundary maps are isomorphisms, hence is for and for odd, and is zero otherwise.

Remark 2.90. In spite of the fact that homology with coefficients determines homology with other coefficient groups, there are many situations where homology with a suitably chosen coefficient group can provide more information than homology with coefficients.

Remark 2.91. In the next chapter we will see that there is a general algebraic formula expressing homology with arbitrary coefficients in terms of homology with coefficients.

Homology and Fundamental Group

Theorem 2.92. By regarding loops as singular -cycles, we obtain a homomorphism . If is path-connected, then is surjective and has kernel the commutator subgroup of , so induces an isomorphism from the abelianization of onto .

Classical Applications

Proposition 2.93. For an embedding , for all .

Proposition 2.94. For an embedding with , is for and otherwise.

Proof.
Proof. Induction by . Write as the union of hemispheres and intersecting in . The Mayer-Vietoris sequence for and , both of which have trivial reduced homology by the last proposition, then gives isomorphisms .

Corollary 2.95. Jordan curve theorem

  A subspace of homeomorphic to separates it into two components, and these components have the same homology groups as a point.

Corollary 2.96. An embedding must be surjective.

Proof.
Proof. As in the last proof, consider the end of the Mayer-Vietoris sequence For , we have must be empty.

Corollary 2.97. In particular, cannot be embedded in since this would yield a nonsurjective embedding in .

A consequence is that there is no embedding for since this would restrict to an embedding of into .

More generally there is no continuous injection for since this too would give an embedding .

Theorem 2.98. Invariance of Domain

If is an open set in and is an embedding, or more generally just a continuous injection, then the image is an open set in .

Corollary 2.99. If is a compact -manifold and is a connected -manifold, then an embedding must be surjective, hence a homeomorphism.

Proposition 2.100. An odd map , satisfying for all , must have odd degree.

Proposition 2.101. Given a two-sheeted covering space , This is the long exact sequence of homology groups associated to a short exact sequence of chain complexes consisting of short exact sequences of chain groups The map is surjective since singular simplices always lift to , as is simply-connected. Each has in fact precisely two lifts and . Because we are using coefficients, the kernel of is generated by the sums . So if we define to send each to the sum of its two lifts, then the image of is the kernel of . Obviously is injective, so we have the short exact sequence indicated.

Since and commute with boundary maps, we have a short exact sequence of chain complexes, yielding the long exact sequence of homology groups.

The map is a special case of more general transfer homomorphisms considered in the next chapter, so we will refer to the long exact sequence involving the maps as the transfer sequence.

Corollary 2.102. Borsuk-Ulam Theorem

For every map there exists a point with .

3Cohomology

Definition 3.1. Consider a chain complex of free abelian groups The dual of is the cochain group , and the dual of is the coboundary map . It follows from that , and we can define the cohomology group .

Definition 3.2. A free resolution of an abelian group is an exact sequence of free groups Appply and we get a chain complex We may lose exactness, and use the temporary notation .

Lemma 3.3. For free resolutions and of , there are canonical isomorphisms for all .

Every has a free resolution of the form . Suppose is generated by , and let be the free abelian group with basis and . Let , which is a subgroup of a free abelian group thus also free, and be the inclusion . For this resolution we get for , which is hence also true for all free resolutions. Thus only is nontrivial, written as .

Theorem 3.4. Universal Coefficient Theorem for Cohomology There is a split exact sequence

Proposition 3.5.

.

if is free.

.

Corollary 3.6. If and of a chain complex of free abelian groups are finitely generated, with torsion subgroups and , then .

Lemma 3.7. If a sequence of abelian groups is exact, then so is .

Lemma 3.8. For free resolutions and of , there are canonical isomorphisms for all .

The group , which depends only on and , is denoted . Since a free resolution always exists, only is nontrivial, and is written as .

Theorem 3.9. Universal Coefficient Theorem for Homology There is a natural exact sequence that splits though not naturally

Proposition 3.10.

.

.

if or is free, or more generally torsionfree.

where is the torsion subgroup of .

.

Corollary 3.11. For finitely generated and , .

Corollary 3.12. (a) , so when , it equals .
(b) If and are finitely gererated, then for prime, consists a summand for: each summand of , each summand of , each summand of .

Remark 3.13. Here is a similar result given in Exercise 2.2.43:
If is a CW complex with finitely many cells in each dimension, then is the direct sum of: a for each summand of , a for each summand of , a for each summand of .

Definition 3.14. A local orientation of at is a choice of generator of , and an orientation of is a function assigning to each a local orientation satisfying: each has a neighborhood , such that all at are the images of one generator under the natural maps . is orientable if an orientation exists.