4th Year Project Contents Chapter 1: Morse functions on Manifolds .................................................................................................................... 2 1.1 - The Morse Lemma .............................................................................................................................................. 2 1.2 – Gradient-Like Vector Fields .............................................................................................................................. 6 Chapter 2: Handlebodies .............................................................................................................................................. 11 2.1 – Handle Decompositions .................................................................................................................................. 11 2.2 – Sliding Handles ................................................................................................................................................ 17 2.3 – Cancelling Handles .......................................................................................................................................... 19 Chapter 3: Handlebodies, Homology and Stability .................................................................................................. 22 3.1 – Homology and Handlebodies ......................................................................................................................... 22 3.2 – Morse-Smale Dynamics ................................................................................................................................... 26 Chapter 4: Morse-Witten Homology ........................................................................................................................... 29 4.1 – Relative Homology Approach ........................................................................................................................ 29 4.2 – Geometrical Interpretation .............................................................................................................................. 32 4.3 – Examples ............................................................................................................................................................ 36 Chapter 5: Poincaré Duality ......................................................................................................................................... 42 5.1 – Cohomology Groups ........................................................................................................................................ 42 5.2 – How to prove Poincaré Duality ...................................................................................................................... 43 5.2.1 The Attaching maps of a Handlebody and Intersection numbers ........................................................ 44 5.2.2 Using Exactness to determine the Sign of an Intersection: ..................................................................... 46 5.3 – The Proof of Poincaré Duality......................................................................................................................... 47 References ....................................................................................................................................................................... 50 0811265 Ian Vincent Page 1 of 50 Chapter 1: Morse functions on Manifolds We recall the notion of a manifold. In this document, all manifolds considered will be smooth, and may or may not have boundary. We first introduce the concept of Morse functions as follows: Definition 1.1 (Degeneracy of Critical Points) Let be a smooth real-valued function on a manifold . Recall the notion of a critical point; that is a point where ( ) ( ) ( ) ) about . Let ( ) be the Hessian matrix with respect to a local coordinate system ( ). We define the critical point of at with respect to a coordinate system ( to be a ( ) degenerate critical point if . Otherwise, the critical point is said to be non-degenerate. In fact, it is an elementary property that the notion of critical points, and degeneracy is welldefined, that is it does not depend on the choice of local coordinate system; ) and ( ) at a critical point , and Lemma 1.2 Given two coordinate systems ( ( ) let ( ) and be the Hessians of with respect to these coordinate systems respectively. ( ) Then ( ) . ( ) are related by the formula ( ) ( ) ( )( ) ) to ( ) evaluated at where ( ) is the Jacobian Matrix of partial derivatives from ( . Then we have ( ) ( ) ( ) ( ) Since the Jacobian ( ) of the coordinate transformation at has a non-zero determinant and ( ) ( ), the above equation implies that ( ) ( ) and hence Proof Observe that degeneracy of ( ) and is independent of choice of local coordinate system. Definition 1.3 (Morse Function) of is non-degenerate. A function is a Morse function if every critical point Remark Usually, the critical points of Morse functions are required by definition to have distinct critical values. I do not assume that here in the definition, and instead in Theorem 1.14, we know that for any Morse function (as defined here) there exists another Morse function ‘arbitrarily close by’ with the same critical points but distinct critical values. The precise notion of ‘arbitrarily )-close and will be discussed later. close’ is ( 1.1 - The Morse Lemma In short, the Morse Lemma states that for a Morse function , there is a local coordinate system about any critical point so that locally, can be written in a standard form. This is extremely useful, and immediately has some important corollaries. 0811265 Ian Vincent Page 2 of 50 Theorem 1.4 (Morse Lemma for dimension ) Let be a non-degenerate critical point of ) about . Then we can choose a local coordinate system ( such that the coordinate representation of with respect to these coordinates has the following standard form ) and is a where corresponds to the origin ( ( ). constant with Remark: First, we recall Sylvester’s Law from Linear Algebra: Sylvester’s Law Any symmetric real matrix A has a diagonalisation of the form where is a diagonal matrix containing the eigenvalues of , and is an orthonormal square matrix containing the eigenvectors. The matrix can be written where is diagonal with entries or matrix transforms to . , and √| is diagonal with |. Then the The number in the statement of Theorem 1.4 corresponds to the number of negative diagonal entries in the Hessian ( ) after diagonalisation. As ( ) is a symmetric matrix, we apply Sylvester’s Law to see that does not depend on the way the Hessian is diagonalised; that is is determined purely by the function and the critical point . Definition 1.5 (Index of a Critical Point) The number in the statement of Theorem 1.4 is called the index of a non-degenerate critical point . It should be noted that the index is an integer between 0 and . ) at the critical point Proof of 1.4 Choose a local coordinate system ( , where ( ) ( ) corresponds to the origin . We may further assume that , replacing by ( ) if necessary. Lemma 1.6 There exist real valued functions the origin so that, ( ) defined in a neighbourhood of ( ∑ ) in some neighbourhood of the origin and such that ( ) ( ) Proof This is a fundamental fact of calculus of several variables; for a proof see any standard textbook on PDEs or multi-variable calculus. Since the point ( ) is a critical point, ( ) Therefore applying Lemma 1.6 again to each of the we can find a neighbourhood of the origin) ( ) ( ) ( such that ( ) ∑ ( ( ) for each smooth functions (defined in ) ) in a neighbourhood of the origin. Therefore 0811265 Ian Vincent . Page 3 of 50 ( Setting ) we get ( ) ) ( ∑ ) (Schwarz 1993) ( and ( ∑ ) ( ( ∑ ) ( ) ). We shall refer to the representation of ( a “representation by a quadratic form” of . ) in ( ) as a “quadratic form representation”, or The idea of the proof is to change the representation of into the quadratic representation in the standard form mentioned in the statement of the theorem. We proceed by induction on the number of terms involved in the quadratic form representing . Let us now compute second order partial derivatives of ( ( If we assume that the critical point assume that ) ) at the origin. We obtain ( ) ( is non-degenerate and so ( ) ) after a suitable linear transformation of the local coordinate system ( ( ) . Since the function is continuous, we see that neighbourhood of the origin. We introduce a new coordinate system ( ), where we define |( √| then we may ∑ ). Then we see is non-zero in a by ) We can easily compute and show that the determinant of the Jacobian of the transformation from ( ) to ( ) at the origin is not zero, so that ( ) is certainly a local coordinate system. The square of is computed as follows: (∑ ∑ | |( ∑ ) (∑ ∑ { Comparing the above with the quadratic form representation ( ∑ ∑ { 0811265 Ian Vincent (∑ (∑ ) ) ) if ) if ) of , we see that if ( ) if Page 4 of 50 In the above, we see that the second terms and thereafter are a sum over so that this part is simplified to a quadratic form representation of fewer variables than ( ). Therefore repeating this process, we proceed by induction on the number of variables to prove that we can represent in a standard form. Then by permuting the as necessary to get all the minuses at the beginning, we obtain the required result. From the Morse Lemma, Theorem 1.4, we easily obtain the following: Corollary 1.7 A non-degenerate critical point is isolated; that is for a non-degenerate critical point of , there exists a neighbourhood of , where no other critical points of lie inside . )) so that Proof Take a critical point . By the Morse Lemma, we can find a chart ( ( { } so that in the in the local coordinate system, there exists has standard form In its standard form, we see immediately that the only critical point of the origin 0, corresponding to the point . in the neighbourhood is Corollary 1.8 A Morse function defined on a compact manifold admits only finitely many critical points. Proof We proceed by deriving a contradiction as follows: assume a Morse function on a compact manifold has infinitely many critical points . By compactness, there is a convergent subsequence ( ) ( ) of this sequence. Let be its limit point. Consider a local coordinate system ( ) defined on a neighbourhood converges to the point subsequence { Since ( is () well, since ( . Since the above subsequence { ( )} , by choosing a further subsequence if necessary, we may assume that the is contained the neighbourhood , its partial derivatives ) and Therefore ( )} of ( )) of . ( ) depend smoothly on the point . The derivatives take the value zero and hence these derivatives take the value zero at is the limit point of as ( ). is a critical point of the function . All the critical points of a Morse Function are non- degenerate, and so by Corollary 1.8, they are isolated. However, the sequence { critical points converging to a critical point which is a contradiction with there can only be finitely many critical points. ( )} consists of isolated. Therefore Of course, the next theorem is extremely important for the rest of this document: Theorem 1.9 (Existence of Morse Functions) Let be a closed -manifold and let be a smooth function defined on . Then there exists a Morse function arbitrarily close to . [The precise nature of this ‘close-ness’ is explained below.] The proof of this theorem is long, beyond the scope of this document and hence omitted. I refer the reader to (Matsumoto 2002), whose proof is relatively easy to follow. Remark We make precise the statement Theorem 1.9: 0811265 Ian Vincent and are “arbitrarily close” mentioned in Page 5 of 50 { } of each equipped with local coordinates, For every , there exists an open cover such that for each the following inequalities hold for all : If ( | ( ) | | ( )| ( ) ( )| ( ) ( )| and are two functions satisfiying the above inequalities for a particular ) are ( )-close. , we say that 1.2 – Gradient-Like Vector Fields We recall the notion of gradient vector field: Definition 1.10 (Gradient Vector Field) Let be a function defined in a coordinate ) be its coordinate system. The gradient vector field for in is neighbourhood and let ( the vector field defined by: Note that for a Morse function in standard form, its gradient vector field is written as: As the gradient vector field is only defined in specific coordinate neighbourhoods, the idea of a Gradient-Like vector field is to globalise the notion of gradient vector field to the entire manifold. Let be a Morse function defined on a closed -manifold. In the following discussion, we always ( ). assume that is a smooth vector field on , possibly denoted by For a Morse function Due to Corollary 1.8, if If ∑ , for , we may use the notation is compact then is a finite set. ( ), by we mean for the set of critical points of ∑ Definition 1.11 (Gradient-like Vector Field) We say that Morse function if the following two conditions hold: Ian Vincent . ( ) is a gradient-like vector field for a 1. away from the critical points of 2. If is a critical point of with index , then has a neighbourhood ) such that has the standard form coordinate system ( ( ) and can be written as its gradient vector field: 0811265 on with a suitable Page 6 of 50 Geometrically, the first condition says that away from the critical points of , direction into which is increasing. points in the Naturally, we need: Theorem 1.12 Suppose that exists a gradient-like vector field is a Morse Function on a compact manifold for . . Then there The proof is beyond the scope of the document and so omitted. The strategy is; given an open cover of , define a vector field on each by the gradient vector field and then find ∑ associated -bump functions . Then after dealing with technicalities, the vector field is a gradient-like vector field for . See (Matsumoto 2002) once more for details. Remark To avoid the business of dealing with -bump functions, it is possible to define the Gradient-Like Vector field in a coordinate-free way as follows, at the expense of knowing the existence of a Riemannian Metric on the manifold: Alternative Definition (Coordinate-free Gradient-like Vector field) Given a manifold , equipped with a Riemannian Metric 〈 〉, for this determines a map 〈 〉 mapping [This map is a linear injective map (by definition of the metric 〈 〉. Moreover, since and are two vector spaces of the same finite dimension, we conclude that is an isomorphism]. We then define the gradient-like vector field ( ) to be the unique such that 〈 〉 , the function mapping any vector to the directional derivative of at in the direction . We now use Gradient-like Vector fields as a useful tool to prove statements to do with Classical ] be a real interval. Use the notation Morse Theory. Let be a Morse function and let [ { }. ( ) [ ] Theorem 1.13 If has no critical value in the interval [ ( ) [ ] product ], then [ ] is diffeomorphic to the Proof Let be a gradient-like vector field for . As away from critical points for , we {critical points of } , which is an open subset of define a new vector field on by Since by hypothesis, [ ] contains no critical points of , it is in the domain of the vector field . Consider the integral curve ( ) of ( ) . Using the which starts at a point of definition of velocity vector, we obtain: ( ( )) Thus the curve 0811265 Ian Vincent ( ) ( ) ( ) ( ) ( ( )) maintains constant speed 1 and Page 7 of 50 ( ) travels “upward” as defined by the value of . Since it starts at the level ( ) [ will reach the level at the time . Define a map ( ) ( ) We can show that is a diffeomorphism by using the facts that at time ] [ ] , it by ( ) depends smoothly on both and and the two distinct integral curves do not meet (by uniqueness of integral curves). ] as follows: given a point ( ) [ Define a map [ ] [ ] we know from the existence and uniqueness of integral curves that there is a unique integral curve of on [ ] [ ]. Let ( ). Then ( ) and define ( ) passing through at some time ( ). This map is smooth since the integral curve depends smoothly on the point and hence and both depend smoothly on . Note that ( ) ( ( )) ( ) and ( ) ( ) ( ) so and hence is a diffeomorphism. ], and with the observation that ( ) [ Therefore we have proved that [ ] [ ] ( ) [ ], we complete the proof of the theorem. ( ) Notice in the Statement of 1.9, that in some sense, Morse functions are ‘dense’ that is ‘in every neighbourhood of any function, one encounters Morse functions’. With this notion, we may in fact assume (with no loss in generality) that given a Morse function , every critical point has a distinct critical value: Theorem 1.14 Let be a Morse function on , and let be its critical points. ̃ Then there exists a Morse function whose critical points have the property ̃( ) ̃( ) for )-close. . Moreover, for any , we can find such an ̃, so that and ̃ are ( Proof We use a similar argument here as hinted in the proof of the existence of a Morse function (Theorem 1.12). ( ) We assume that the critical value of at the critical points and are the same, ( ) and try to modify slightly. By the Morse Lemma we can choose a local coordinate system ( ) about and write in the standard form Let be a gradient vector field for ( ) with respect to this coordinate system. Then ( ) ( ) ( ) For a sufficiently small , consider the (closed) -discs of radii and and their respective images and in the local coordinate system above, centred at . It follows from the above ( ) n the region equality that int ( ) Denote by the compact set and by the interior int , so that is an open set containing a ). We extend to a compact set . Consider a -bump function with respect to ( smooth function on the entire manifold by setting outside , and denote the extended ̃ function again (for convenience) by . Define a function by ̃ where is a small 0811265 Ian Vincent Page 8 of 50 ̃ outside , the critical points of and ̃ coincide there. Similarly, since on the interior int( ), we see that the point is the only critical point of both and ̃ in this region. real number. Since Therefore the only place where ̃ possibly has a different set of critical points from between As is the region . We compute differences of the first partial derivatives of and ̃ ̃ | | | | and is bounded for each by the Extreme Value theorem (applied to the compact region follows that by making ), it arbitrary small, the difference |∑ ( ) ∑( ̃ ) | is made arbitrarily small. Since ∑ ( ) between ∑( and so ̃ has no critical points between and ̃ and , by taking sufficiently small, then ) , and so and ̃ have the same critical points. Observe that the non-degenerate critical points for and ̃ also coincide, we see that ̃ is a Morse ̃( ) ( ) ( ) therefore, even though ( ) function. Furthermore, we have ̃( ) ̃( ). ( ), ̃( ) By repeating the above argument as necessary for pairs of critical points whose critical values coincide, (since there are only finitely many critical points for the Morse function ) we obtain that ̃ is a Morse function with distinct critical values. To prove that ̃ is ( Again, since )-close to , we see that ̃ ( ) ( )| | | is bounded on the compact set the difference in the second derivatives of ( )| , by making arbitrarily small we can make and ̃ arbitrarily small. We finish the chapter with the following useful Lemma: Lemma 1.15 (Product of two Morse functions) For compact manifolds and with and , let and be Morse functions. Define a function by the formula ( )( ) where and are positive numbers. Then for and sufficiently large, is a Morse function ) whose critical points are of the form ( where is a critical point for and is a critical point for ; its index is equal to the sum of the indices for and . Proof Let ( about and 0811265 ) be a point of and let ( respectively. Then we can take ( Ian Vincent ) and ( ) be local coordinate systems ) as a local coordinate system Page 9 of 50 about ( ) in . Observe that )( ( ) and ( )( ) Since and are compact, by the Extreme Value Theorem, and are bounded and so there | | and | | . Then the exists and large enough in and respectively so that following two conditions become equivalent: 1. ( 2. ( ) ) ( ) for all for all and ( ) and for all . is of the form ( From the above we immediately have that a critical point of points and for and respectively, and that )( ( Therefore the Hessian of ) and ( )( ) for critical ) is of the form ( ) ( ( ) ( ) ( ( ) ( )( ) ( ) ( ) and so Therefore since and are Morse functions, by definition )( ) have chosen and sufficiently large so that ( so is a Morse Function. ) ) ( ) ( ) . Therefore ( ( ) and we ) , and ) are the local Applying the Morse Lemma to , we may assume that ( ( ) ) and coordinates about to express in the standard form. It follows that ( ( ) are the local coordinates about and to express and in the standard form, and so ) is equal to the sum of the indices of and by equating both sides, we see that the index of ( . 0811265 Ian Vincent Page 10 of 50 Chapter 2: Handlebodies In this chapter, we use Morse functions and gradient-like vector fields to decompose manifolds into objects called handlebodies. This technique is the main focus of Classical Morse Theory. 2.1 – Handle Decompositions During this chapter, we will meet the idea of ‘gluing’ two manifolds with boundary along their boundaries. This leads to some technical issues where the gluing must be done in such a way to preserve the smooth nature of the resulting manifold. The following theorem makes precise the process: Theorem 2.1 (Gluing Manifolds with boundary) Let and be manifolds with boundary, and let be a diffeomorphism between the boundaries. Then we can construct a new manifold by identifying each point with the point ( ) . The resulting manifold is unique up to diffeomorphism. (It is permissible to only identify a few of the connected components of the boundary, instead of the entire boundary.) See the diagram below: The next theorem is more delicate that its predecessor, which says that this gluing is preserved under diffeomorphisms of each component: Theorem 2.2 (Gluing Diffeomorphisms) Let and be the manifolds obtained by gluing manifolds with boundary (where and are diffeomorphisms). Suppose that we have diffeomorphisms and such that ( ) ( ) for every point in . Then there exists a diffeomorphism obtained by identifying and along the boundary. ( ) so Note that the identifications form an equivalence class: ( ) ( ) equivalence classes in its image; that is ( ( )) 0811265 Ian Vincent must respect Page 11 of 50 I do not wish to go into detail with the proofs of these theorems. Intuitively, I think it is clear that the process can be done, but in order to construct an explicit formula for the diffeomorphism in theorem 2.2 is a long-winded process. Regardless, for a proof, see (Milnor 1965) page 25. Now with some technicalities dealt with, we can proceed with the first theorem, which is the fundamental concept of handle attachments: Let be a closed manifold and a Morse function on . We use the notation { } ( ) for a value in the image of . We will investigate how changes at the parameter is varied. Theorem 2.3 If and write has no critical values in the interval [ . ] then and are diffeomorphic, Proof Since is compact and is smooth, the minimum and maximum values for , and are obviously critical values and so the interval [ ] does not contain either or . We may therefore prove Theorem 2.3 under the assumption . { ( ) As before, denote the set [ ] assumption [ ] contains no critical points of . }. Clearly we have [ ] . By By Corollary 1.8, has only a finite number of critical points. We may therefore assume that has no critical points in [ ] for a small enough positive number . By Theorem 1.13, [ ] in this case. We also note that [ [ ] is diffeomorphic to the product ] has no critical points in [ [ ] so that ] and hence by Theorem 1.13, [ ] is [ ] diffeomorphic to the product . Therefore we have a diffeomorphism [ where we may assume that the restriction of ] [ ] to the level set is the identity map. ). We now define another diffeomorphism -bump function with respect to ( to “glue together” the [ ] [ ] using Theorem 2.2 and diffeomorphism on [ to get the required diffeomorphism ] and the identity map on . (I hope the idea is clear; Theorem 2.2 guarantees this will work, but to form an explicit formula is complicated and provides little in the way of clarity). Let be a From the above theorem, the problem we need to investigate is the change in shape of has the parameter passes through a critical value. From Theorem 1.14, we may assume that takes distinct critical values at distinct critical points. We also notice that has a finite number, say , of critical points. By permuting the critical values of so that they lie in ascending order, we can label their corresponding critical points as ; so ( ) . ( ) whenever Let ( ), the critical value of when . We begin by analysing the changes of . It is immediate that around and when , and similarly, : By Theorem 1.14, we may assume that is the only point which gives the minimum value. Using the Morse Lemma, we can write in the standard form . Observe that the index of is necessarily zero, because is the minimum of on . 0811265 Ian Vincent Page 12 of 50 Let be a sufficiently small number (for example, so {( the Morse Lemma we can express ) ). Clearly, but from } ; that is is diffeomorphic to the -disc . The standard form shows that takes the minimum value the centre of the disc, and attains its maximum on the boundary of the disc. Observe that this argument can be repeated whenever disc so is diffeomorphic to the disjoint union critical point of index is called an at is a critical value of index 0; we add an . The -disc appearing at each -dimensional -handle, or -handle for short. Similarly, at the maximum value , by Theorem 1.14 we may assume that is the only point ( ) where . It follows immediately from the Morse Lemma and maximality of that in standard form, and hence the index of is necessarily . Using these ) local coordinates we see {( , which corresponds to the complement of the diffeomorphic to -disc of radius √ in . In this case, the boundary is . As increases from and passes , we see that the boundary of is capped off with an -disc, and forms the completion of a compact manifold without boundary. More generally, whenever passes any critical value of index , the -handle caps off a connected component of the boundary ; this component is diffeomorphic to . We now have a complete idea of what happens around critical points of indices and . Next we consider the changes of near a critical point of a general index as passes through the corresponding critical value . First, we use the Morse Lemma to take a local coordinate system about the critical point have in the standard form: . The situation will around idea: Around 0811265 will be clarified in Theorem 2.5. The image below shows the general , the darkly shaded area in the above diagram corresponds to setting ; that is the inequalities : Ian Vincent and which we get by . The lightly shaded area corresponds to Page 13 of 50 { where is another positive number much smaller than . This lightly shaded area is called an dimensional handle of index , or more briefly, an -dimensional -handle. This handle is diffeomorphic to the direct product . Definition 2.4 (The core and cocore of a -handle) ) ∑ {( The -disc } is called the core of the -handle, and the ( ) ∑ {( )-disc } intersecting the core is called the cocore. The name cocore indicates the thickness of the -handle. The core and cocore intersect transversely at the origin, which is exactly the critical point . See Definition 2.10 later for the concept of ‘transverse intersection’. We attach a -handle to as shown by the thick lines in the diagram above. Then ( is diffeomorphic to the union Theorem 2.5 The set . That is Strictly speaking, the space ). is diffeomorphic to the manifold obtained by attaching a -handle to ( ). with a -handle attached is not “smooth” at the “corners” of the boundary where the handle meets . We must “smooth” out these corners to take a manifold as shown in the diagram below. A more accurate statement for Theorem 2.5 therefore, should be that is diffeomorphic to the smoothed-out manifold . Notice that in the diagram, corresponds to the set {( ) ∑ ∑ } Idea of Proof of Theorem 2.5. The idea is similar to the proof of Theorem 2.3. We again use a gradient-like vector field for . One can see in the diagram above that the vector field , after 0811265 Ian Vincent Page 14 of 50 leaving the boundary reaches the boundary let flow along and We of of continues to flow upward (that is, with respect to ) until it . We may multiply the vector field by a suitable function, and so that it will match after a certain period of time. This will show that are diffeomorphic. will not worry too much about the non-smooth edges of the boundary , and we will pretend that it is in fact a smoothed-out manifold in our discussion. By looking at the standard form of at and the cocore is diffeomorphic to a ( , we see that the core )-disc. is diffeomorphic to a” -disc, The core of a -handle is -dimensional and the cocore is -dimensional, so that there is no downward direction; every direction points upward. On the other hand, the core of an -handle is -dimensional so that any -handle “faces down”. Definition 2.6 (The attaching map of a -handle.) One attaches a -handle pasting along the boundary to by (indicated by the thickened lines in the above diagram). In order to describe the handle-attaching accurately, one must specify a map indicating where each point of corresponds to in the boundary . The map is a smooth embedding which we call the attaching map of the -handle. The boundary of the core )-dimensional sphere disc is a ( , and is called the attaching sphere. [Note that in the case , we find that and define instead the “attaching map” to be the disjoint union .] An attaching map is an embedding into the boundary of a thickened ( )-sphere . Example 2.7 (A 3-dimensional 1-handle and a 2-handle). In the diagrams below, a 3-dimensional 1-handle and a 2-handle are depicted, respectively. The picture of the 1-handle justifies the name “handle”. The 2-handle is depicted as a thickened upside-down “bowl”. 0811265 Ian Vincent Page 15 of 50 Definition 2.7 (Handlebody) A manifold (with boundary in general) obtained from by attaching handles of various indices one after another ( ) ( ) is called an dimensional handlebody. More precisely, a handlebody is defined in three steps as follows: 1. A disc is an -dimensional handlebody. 2. The manifold attaching map of class ). by ( ( 3. If ( ) obtained from , ) by attaching a , is an is an -dimensional ( -handle with an -dimensional handlebody, denoted handlebody, then the manifold ) obtained from by attaching a -handle with an attaching map ) of class , is an -dimensional handlebody, denoted by ( [Strictly speaking, the manifold is “smoothed out” each time a handle is attached, so that the resulting handlebodies are always considered smooth manifolds.] Remark It is worth noting that the attachment of a -handle is nothing but a disjoint union so in the case of index , there is no need to specify the attaching map . Therefore in this case, the ) does not have meaning by itself, but in the attaching map in the notation ( notation is traditionally kept as a formality. Theorem 2.8 (Handle decomposition of a manifold) When a Morse function is given on a closed manifold , a structure of a handlebody on is determined by . The handles of the handlebody correspond to the critical points of , and the indices of the handles coincide with the indices of the corresponding critical points. In other words, can be expressed as a handlebody. When a manifold is expressed as a handlebody, it is called a handle decomposition. Proof We may assume that all the critical points of the given Morse function have distinct critical values. Permute the critical points in such a way that their critical values are in ascending order and name them . Let be the index of the critical point . Fix a gradient-like vector field on for . The proof now proceeds by induction on the subscripts of the critical points . Let be the value of at , and we will show that is a handlebody. First, for , the index of the critical point is , as gives the minimum value of . Therefore, is diffeomorphic to the -dimensional disc . From part 1 of Definition 2.7, is indeed an -dimensional handlebody, and the statement for the induction is proved for . In this case, is a -handle itself. Next, we make the inductive assumption that prove that is a handlebody ( diffeomorphic to a manifold obtained by attaching a 0811265 Ian Vincent ) and we will is a handlebody ( ) . Recall from Theorem 2.5 that is -handle. The attaching map Page 16 of 50 of this handle is determined naturally without ambiguity from the discussion following Theorem 2.5. ] contains no critical values, so from Theorem 2.3, The interval [ to . By consulting the proof, this diffeomorphism is given by letting gradient-like vector field until it matches From the induction hypothesis that . Let is a handlebody be such a diffeomorphism. ( diffeomorphic to the same handlebody. Therefore the manifold attaching a is diffeomorphic flow along the ), we see that , which is obtained from is by -handle is also diffeomorphic to a handlebody from part 3 of Definition 2.7. This completes the proof of Theorem 2.8. However, since it is important, let’s investigate the attaching map of the new -handle in more detail: ) so that the above From the induction hypothesis, is the handlebody ( ) to diffeomorphism can be seen as a diffeomorphism from ( . Precisely ), but rather is -handle is not attached directly to the handlebody ( and its attaching map is naturally determined. Therefore, ) by the diffeomorphism , the is identified with the handlebody ( ( ) by the composition map is attached to the handlebody { ( )} If we denote this composition by , then is speaking, the attached to when handle indeed the handlebody ( ) ( ) Remark The above proof implies the following: when a handle decomposition of a manifold from a Morse function , is obtained 1. The order of the handles and their indices are determined by the critical points of , and )of the handles are determined by a gradient-like vector field 2. The attaching maps ( for (since is determined by ). 2.2 – Sliding Handles In fact, there is some flexibility to the attaching maps of handlebodies; intuitively, ‘sliding the handles around’ does not change the diffeomorphism type of the handlebody. By ‘sliding’ we actually mean composing the attaching map of a -handle with an isotopy of the boundary of the subhandlebody formed by all the previous handle attachments. First we define ‘Isotopy’, ‘Transversal Intersection’ and make precise the notion of ‘General Position’: Definition 2.9 (Isotopy) Let be a -dimensional manifold. The set { } is called a family of diffeomorphisms if, for each real number in the open interval , a diffeomorphism is assigned. The family { } is called an isotopy of if the following two conditions are satisfied: 1. The open interval contains the closed interval [ ], and is constantly the identity map on for . Also, for all , we have where is some diffeomorphism of . 2. The map defined by ( ) ( ( ) ) is a diffeomorphism. That is the map depends on the parameter smoothly in this sense. 0811265 Ian Vincent Page 17 of 50 Definition 2.10 (Transversal Intersection) Suppose that there are an -dimensional manifold and a -dimensional manifold in a -dimensional manifold such that . We say that and intersect transversely at a point of , and write , if for every point the tangent spaces and have the property . In particular, for two transverse submanifolds ( ) Lemma 2.11 (General Position) Let and respectively, in a -dimensional manifold . I. II. If If and , we have ( ) ( ) be compact submanifolds of dimensions , then there is an isotopy { } of such that then there is an isotopy { } of such that transversely in finitely many points. and and ( ) . and ( ) intersects Proof Omitted. See (Milnor 1965), pages 46, 47 for a Proof. Theorem 2.12 (Sliding Handles) attaching map further that an isotopy { } of Suppose that a -handle is attached by an on an -dimensional manifold with boundary, and suppose is given. Here, we assume that and . Then the new handlebody of the handle by the isotopy, is diffeomorphic to the original handlebody (before the -handle was ‘slid’). Notice that the only difference between the two handlebodies is that the attaching map has been replaced by another attaching map , where is some diffeomorphism. In other words, the handle has been deformed so that it lies somewhere else on the boundary . The idea is given in the diagram to the right: Proof The proof is long and not especially enlightening. Therefore it is omitted. From the theorem above, we have the following Corollary, which is the main result of this section: Corollary 2.13 (Rearrangement of Handles) Any handlebody can be modified in such a way that the new one is described as follows: It is constructed first from a disjoint union of -handles, and then a disjoint union of 1-handles are attached to them, and then a disjoint union of 2-handles are attached, and so forth, so that handles are attached in ascending order of indices. Rearranging the handlebody in this manner is equivalent to modifying the original Morse function (from which the original handlebody is constructed) via compositions of isotopies to another Morse function (from which the modified handlebody with these required properties is constructed.) The diagram below illustrates the situation of the critical points of the Morse function . As we see, is obtained from the handlebody which consists of handles of indices less than or equal to all attached, by attaching a disjoint union of copies of -handles 0811265 Ian Vincent Page 18 of 50 at the same time. Here, is the number of handles of index . Proof (Sketch) The proof is a sequence of technical Lemmas which all involve performing isotopies to the gradient-like vector field for . The argument is too long to contain here, and I refer the reader to (Matsumoto 2002), pages 105-120. As a sketch, the idea is to is to use Theorem 2.12 repeatedly to slide the handles into a particular order: Let be the sub-handlebody consisting of handles of index at most . The goal is by sliding, we ensure that for every , every -handle is attached to and all the attachments of handles to happen disjointly (that is, we slide the -handles off each other). Note that here; whenever I said ‘sliding’, what I actually meant is ‘composing the attaching maps with an isotopy’, but in my opinion greater understanding is achieved by thinking of it as ‘sliding’. 2.3 – Cancelling Handles In this section, we describe a situation where two consecutive handle attachments can be ignored; that is after attaching two consecutive handles of specific indices under a specific family of attaching maps, the result is diffeomorphic to the original handlebody. Definition 2.14 (Belt Sphere) Consider the situation where a -handle is attached to an dimensional manifold with boundary. Set . In this case, the boundary of the cocore of this -handle. By definition, the belt sphere is an ( as in the diagram below: - of the -handle is called the belt sphere )-dimensional sphere embedded in Theorem 2.15 (Cancelling Handles) Suppose that a manifold is obtained from an dimensional manifold with boundary by attaching a -handle, and suppose further that a )-handle: manifold is obtained from by attaching a ( 0811265 Ian Vincent Page 19 of 50 If the belt sphere of the -handle and the attaching sphere handle intersect transversely at a single point in the boundary of , then to . )of the ( is diffeomorphic Proof Omitted, see (Matsumoto 2002) pages 120-125. The statement of Theorem 2.15 may seem more reasonable after looking at some examples: 1. The first diagram below illustrates the case where a -handle and a 1-handle are attached to . In this case, the belt sphere of the -handle is a 2-dimensional sphere which is the surface of a 3-dimensional disc (shown in blue) and the attaching sphere of the 1-handle is two points (shown in red), one of which lies on the surface of the 0-handle. As we see from the diagram, the union of the 0-handle and the 1-handle can be squashed into . 2. Next, the diagram below illustrates the situation where a 1-handle and a 2-handle are attached. The attaching sphere (red) of the 2-handle and the belt sphere (blue) of the 1handle intersect transversely at a single point. In this case, the union of the 1-handle and the 2-handle can be swallowed into as well. From the theorems above to do with sliding and cancelling handles, we have the following theorem, allowing us to assume that there is only one critical point of index and only one of index : Theorem 2.16 Let be a closed -dimensional manifold. If is connected, then there is a Morse function on with only one critical point of index and one critical point of index . (The Morse function may have as many other critical points of other indices with ). Proof By Corollary 2.13, for some Morse function , we can assume all the critical points of index 0 take the same critical value , all the critical points of index 1 take the same critical ) and so forth. value ( 0811265 Ian Vincent Page 20 of 50 The handle decomposition with respect to such a Morse function is of the form: and If (a disjoint union of handles) by attaching some -handles. is obtained from is not connected, then it consists of more than one connected component then, since we attach handles of indices 2 or higher to obtain from , but , and the attaching sphere of a handle of index 2 or higher is connected (it is diffeomorphic to ), it must be attached to a single connected component. Hence the number of connected components does not increase or decrease after attaching handles of index 2 or higher, and hence if is disconnected, then is disconnected. To avoid a contradiction, we must have connected. Suppose the handle decomposition of has more than one connected component. From the above consideration, they must all be “bridged” together and become connected as a whole after attaching the 1-handles. One among the expression above is bridged to another by a 1( handle . Since the attaching sphere points) of this one handle intersects the belt sphere of the 0-handle at exactly one point, by Theorem 2.15 (Cancelling Handles), the critical point corresponding to the 0-handle and the critical point corresponding to the 1handle are cancelled out together. Repeating the argument, we obtain a Morse function with only one critical point of index . Next, we consider the function with the sign reversed, function obtained from by setting (– )( ) has index ( ) where and – , the index of a critical point sets of critical points are the same for critical point of – , . The function – is the Morse ( ). Although the of is , then as a . In terms of handle decompositions, the roles of “core” and “cocore” are interchanged, and a handle of becomes an ( )-handle of – . Using the same argument as above and Theorem 2.15, we perturb – and reduce the number of handles to 1. Then it can be considered that the number of -handles of was reduced to . In this case, the number of -handles of – (that is, the number of -handles of ) stays unchanged and is 1, so that the perturbed has only one 0-handle and one -handle. 0811265 Ian Vincent Page 21 of 50 Chapter 3: Handlebodies, Homology and Stability 3.1 – Homology and Handlebodies In the previous chapter, we saw how to decompose a closed manifold into a series of handle attachments. First, we must clarify the relationship between this handle decomposition (which gives rise to a cell decomposition of the manifold) and the homology of the manifold. The idea is that the handle decomposition is akin to dividing up a manifold into a cell complex, which we then use to compute the homology of the manifold. In order to prove the theorem, we must first define the notion of Mapping Cylinder: Definition 3.1 (Mapping Cylinder) For a continuous map between topological spaces, [ ], [ ], the space obtained from by attaching by identifying the point “at the [ ] and the point ( ) of for each point bottom” ( ) of the direct product is called the mapping cylinder of , and is denoted by . Lemma 3.2 The mapping cylinder of the inclusion is a homotopy equivalence. Proof We construct a continuous map is homotopy equivalent to . More precisely, as follows. [ ], we set ( ) Looking at the structure for a point of , and set ( ) [ ]. Then it is clear that a continuous map ( ) for a point ( ) of is defined. We now show that and are homotopy inverses of each other. It is clear that . To show , we construct a homotopy [ ] from to as follows. [ ]. ) for a point ( ) of Namely, set ( ) for a point of and set (( ) ) ( [ ], and hence, a ) is a point of In this expression is a parameter of the homotopy and ( point of . When , we easily see that have shown . gives the identity map on and gives when , so we [ ] in In short, is a continuous deformation which collapses the product part down to { } gradually. Hence gives a homotopy equivalence , and also we see that and are homotopy inverses of each other. It is best to see Lemma 3.2 in action with a simple example: 0811265 Ian Vincent Page 22 of 50 Example 3.3 The -disc is homeomorphic to the mapping cylinder of the map { }, which collapses the ( )-sphere to a single point . By Lemma 3.2, is homotopy equivalent to { } so is homotopy equivalent to a point { }. [Of course, this example is so easy the application of Lemma 3.2 seems unnecessary; one could easily prove is homotopy equivalent to a point without it] The main result of this section is the following theorem: Theorem 3.4 (Handlebodies and Cell Complexes) Let be an -dimensional handlebody. If the largest index of the handles contained in is , then is homotopy equivalent to an -dimensional cell complex . More precisely: 1. There exists a continuous map from the boundary of to such that is homeomorphic to the mapping cylinder of . Therefore by Lemma 3.2, . 2. There is a one-to-one correspondence between the -handles of and the -cells of . Proof The idea is as follows: we reduce the radii of the cocores of handles smaller and smaller, and finally shrink the handles to their cores. The cell complex is one obtained by attaching the discs of cores of handles one after another. By Corollary 2.13 (Rearrangement of Handles), we can assume that following form: ( where ) represents a -handle Specifically, ( ) . The notation ( is a handlebody of the ) denotes the th -handle. is constructed from a disjoint union of -handles by attaching a disjoint union of -handles and so on. Theorem 3.4 is proved by induction on the maximal index of handles contained in . If , then , copies of -discs. If we regard the set of points { } as a dimensional cell complex , then by Example 3.3, is homeomorphic to the mapping cylinder { of the map } which collapses each sphere to a single point. Assume Theorem 3.4 is proved for handlebodies consisting of handles of indices less than or equal to , and prove it for a handlebody whose maximal index of handles contained in is . Let be the subhandlebody of ( form consisting of all handles of indices less than . Thus ) . By the induction hypothesis, there is a cell complex continuous map such that is homeomorphic to the mapping cylinder simplicity, we now assume that there is only one -handle attached to : has the and a of . For The handle can be regarded as a mapping cylinder. Namely, if { } { }, then ) denotes the map which sends any ( to the core ( ) is homeomorphic to the mapping cylinder of . By Example 3.3, regard as a mapping 0811265 Ian Vincent Page 23 of 50 cylinder of the second component by of this mapping cylinder. ; then can also be regarded as the product Our strategy for the proof of Theorem 3.4 is to construct an -dimensional cell complex { } and construct a continuous map and from and . from The attaching map regarded as a submanifold of can be is a smooth embedding, so that via . We make this identification in what follows. { } is also a submanifold of Then, { } with the attaching map . . We denote the restriction of on { } To simplify the notation, we set ( of can be ); then the boundary decomposed as Our desired continuous map { }, where the map we used when the handle is defined on the portion by { } is regarded as an -cell in (attached to ). Also, is is regarded as a mapping cylinder . )-manifold with boundary, and consider the collar To define on , we regard as an ( neighbourhood (here it is a closed collar neighbourhood) of the boundary in . Thus we have [ ] and { } . { } of the map may be identified with with the collar Let be the mapping cylinder of the restriction | which we used to identify the handle with a mapping cylinder. Then , and is homeomorphic to (which is attached from outside). Let ( ) ] to [ be this homeomorphism. Note that by mapping { } [ ], we have a natural map cylinder { }( [ ] ) we have ) we have id and on Now, a desired continuous map ( ) 0811265 Ian Vincent is defined on { [ . On | ] inside the mapping { }( [ ] . as follows: ( ) ( ) ( [ ]) Page 24 of 50 Since the two maps and coincide on the intersection RHS, a continuous map | on is well defined. { } of the two regions in the The map coincides on the intersection . defined on each portion of the decomposition of , so we have obtained the desired continuous map From the construction of , it is clear that is homeomorphic to the mapping cylinder , and that there is a one-to-one correspondence between -cells of the -dimensional cell complex and handles of . Remark The cell complex in the proof of Theorem 3.4 is considered to be embedded in the handlebody from its construction. In particular, if the handlebody is a closed manifold, then and can be identified. For, if then the mapping cylinder of the map is nothing but itself. For example, as we can see the -sphere regarded as an -dimensional cell complex consisting of a single 0-cell (a point) and a single -cell whose boundary is identified to that point. Now we gone to the effort to prove Theorem 3.4, we can use handlebody theory as a link between the homology and Morse functions of the manifold. A classical result of Morse Theory is the following: Theorem 3.5 (Morse Inequality) Let be a closed -manifold, and a Morse function on . For the number of critical points of index and the -dimensional Betti number ( ) ( ( )) of , the following inequality holds: ( ) Proof Let be a Morse Function on a closed of critical points of of index . -manifold . We denote by the number Consider the handle decomposition defined by . Then by Theorem 3.4 (Handlebodies and Cell Complexes) and the above remark, can be identified with a cell complex , and there is a one-toone correspondence between cells contained in and the handles of . In particular, the number of -cells of equals the number of -handles of . Consider the Cellular Chain Complex of ( )]: [here we use the notation as in (Hatcher 2001); ( )→ ( )→ → For each , the rank of ( ) is equal to the number and ( ) ( ) ( ) of -cells of . Denote by ( ) Since the th dimensional homology group ( ) is obtained from a subgroup ( ) ( ) by taking the quotient by a smaller subgroup ( ) ( ), we have ( ) ( ) ( ) by the identification of and , we have ( ) ( ) ( ) ( ) for all . This proves the Morse inequality. and therefore we obtain 0811265 Ian Vincent Page 25 of 50 ( ) Observe that the Betti numbers are determined by uniquely; from this inequality we see that the number of critical points of any Morse function on is restricted by the ‘shape’ of . In particular, if ( ) , then a Morse function on must have at least one critical point of index . 3.2 – Morse-Smale Dynamics We now focus our attention away from handlebodies temporarily and back to Morse functions, or more specifically, to their gradient flows. As usual, let denote by be a Morse function, and let ( ) the flow on integral curve of – ( ) be a gradient-like vector field of . We ( ) is the determined by – , that is for fixed , the curve starting at the point , so ( ) , ( ( )) ( )) . From the ( theorem of existence and uniqueness of solutions to ODEs, we see that ( diffeomorphism on . ) ( ) is a Example 3.6 To familiarise ourselves with the notion, consider the sphere embedded in in the usual way, and be the ) Morse function defined by ( the ‘height’ function. Then is pictured as shown in blue, and ( ) gives, for a given , the point on the integral curve (red) of based at at time . Definition 3.7 (Stable/Unstable Manifolds of a critical point) Suppose ( ). Fix a gradient-like vector field on function and and by the flow of is a Morse and as before denote generated by – . We set ( ) ( ) { ( ) } ( ) is called the stable manifold of ( ) is called the unstable (relative to the and }. ) ( ) { gradient-like vector field ), we set ( The following diagram illustrates the concepts defined above: 0811265 Ian Vincent Page 26 of 50 Lemma 3.8 Let and . For sufficiently small embedded in the level set { be the index of a critical point for a Morse function ) is a sphere of dimension , the set ( smoothly }. Proof Use the Morse Lemma to pick local coordinates {| | sufficiently small so that in the neighbourhood form ∑ In the more concise notation ( | | ) adapted to . Fix }, the vector field has the ∑ . A trajectory, (or flow line) ( ) of – which converges to as must stay inside for all sufficiently large (and negative). Inside , the only such trajectories have the form , and they }. Moreover, since is strictly decreasing ( ) { | | are all included in the disc ( ) ( ) . Being the on nonconstant trajectories, we deduce that if then ) is homeomorphic to boundary of an embedded disc, we conclude that ( . ( ) is a smooth manifold diffeomorphic to Proposition 3.9 smooth manifold diffeomorphic to . Proof We need only prove the statement for the unstable manifold ( ) ( ). like vector field for and Using Lemma 3.8 fix a diffeomorphism coordinates of , we can define ( ( ). If ( , while ( ) is a ( ) since – is a gradient- ) with denote the polar ( ) by ) ( ( )) From the definition of ( ), it is a diffeomorphism. Being the composition of diffeomorphisms, we therefore see that is a diffeomorphism as required. The following Theorem allows us to restrict the theory to just dealing with the case where the unstable manifolds and stable manifolds intersect transversely. Theorem 3.10 Suppose gradient-like vector field for intersects the stable manifold is a Morse function on a compact manifold . Then there exists a ( ) such that for any , , the unstable manifold ( ) transversally. Proof Since it is a lengthy proof, we refer the reader to (Nicolaescu 2007), pgs 57-60. For convenience, we package these above properties into one definition: Defininiton 3.11 (Morse-Smale Pair, Morse-Smale Vector field) If is a Morse function and is a gradient-like vector field for , such that ( ) ( ) Then we say that ( ) is a Morse-Smale pair on and that is a Morse-Smale vector field adapted to . 0811265 Ian Vincent Page 27 of 50 From Corollary 2.13 (Rearrangement of Handles), in the construction of a handlebody from a Morse function, we are allowed to attach the handles in ascending order of their index, and under the conditions of the Lemma, this does not change the diffeomorphism type of the handlebody. In terms of Morse functions, this Lemma suggest that without loss of generality, we may as well ( ) for all assume that the Morse function has the property ( ) whenever ( ) ( ). ( ) of extends ‘downwards’ with respect to and in Intuively then, the unstable manifold ( ) of extends ‘upwards’ particular, ‘away’ from the point . Similarly, the stable manifold with respect to , and away from the point . By imagining the situation, it seems that the unstable and stable manifolds of and respectively should be disjoint. This is precisely what we prove now. If ( ) is a Morse-Smale pair on and ( ), then ( ) ( ) points such that ( ) . Proposition 3.12 ( Proof Suppose not. By Definition 3.11, ( ( ) ( )) ( ( ) and therefore ( ( ) ) intersects ( )) ( ( )) ( . Assuming ( ) ( are two distinct critical ( ) transversally we have ( ( ( )) ( )) ( ) ) ( ) ( ) , then )) ( ) ( ) contains non trivial flow By definition, if non-empty, then the intersection lines; the idea is that these lines “diverge” from and “converge” to , and . In particular, each flow line has dimension 1. This gives a contradiction; the 0-dimensional subspace ( ) ( ) cannot contain a 1-dimensional flow line. Definition 3.13 (Self-indexing Morse function) ( ) for every indexing if ( ) . A Morse function is called self- The following theorem is proved using handlebody decompositions in (Matsumoto 2002). The proof is long, nasty and involves lots of technicalities. In (Nicolaescu 2007), there is a shorter proof based on (Milnor 1965), pages 37-44 using the theory developed in this section. Theorem 3.14 (Smale) Suppose there exist Morse-Smale pairs ( ) on is a compact smooth manifold of dimension such that is self-indexing. . Then In fact, the same proof of Theorem 3.14 as in (Nicolaescu 2007) can be used to prove the stronger (and more useful statement below): Corollary 3.15 Suppose ( ) is a Morse-Smale pair on the compact manifold can modify to a smooth Morse function with the following properties: and ( 1. 2. ( ) ( ) for all is a gradient-like vector field for . In particular, ( 0811265 ) . Then we ) is a self-indexing Morse-Smale Pair. Ian Vincent Page 28 of 50 Chapter 4: Morse-Witten Homology Armed with the Theory of Handlebodies (Chapter 2) and section 2.2 on Morse-Smale Dynamics, we can define the Morse-Witten Homology groups for compact manifolds. We look at two methods. The first is a little more precise and utilises the relative homology of the manifold, while the second is more geometric. 4.1 – Relative Homology Approach In this section, we construct the Morse-Witten Homology using handlebodies. This method of construction is very similar to the construction of Cellular Homology. For this section, suppose that ( ) is a Morse-Smale pair on a compact such that is self-indexing. In particular, note that the real numbers when . We set { } { -dimensional manifold are regular values of } Then is a smooth manifold with boundary which can be decomposed into the ‘upper’ and ‘lower’ boundaries and respectively: { where } Recall the following theorems about computing homology groups (as stated in (Hatcher 2001)): (Excision) Given subspaces ) then the inclusion ( all . Equivalently, for subspaces ) ( isomorphisms ( (Homology of a Wedge Sum) an isomorphism ( such that the closure of is contained in the interior of ) induces isomorphisms ( ) ( ) for , the inclusion ( which cover ) for all . For a wedge sum ⋁ ⨁ ̃( ⨁ ̃ (⋁ ) provided the wedge sum is formed at base points some neighbourhood of in . (Homology of a Quotient) quotient map ( ) ( If , the inclusions ( ) induces ⋁ induce ) , such that for a manifold ) , with is a deformation retract of closed and non-empty the ) induces isomorphisms ( ) ( ) ̃( ) for all (Where ̃ denotes the reduced homology. In fact, for any topological space , . See (Hatcher 2001) for details. ( ) ̃ ( ) for Remark Hatcher deals with a more general version of this statement; he defines a pair ( ) to be good if is closed, non-empty and is a deformation retract of some neighbourhood 0811265 Ian Vincent Page 29 of 50 ) to be a good pair if in . The fact that is a manifold in the above statement guarantees ( with closed non-empty; take an open neighbourhood of consisting of a union arbitrarily small open balls centred about the each point of , then define . Then is clearly a deformation retract of , since a ball is a deformation retract of a point, and hence is a deformation retract of . of ( ) We define ( ), the th relative homology group of the pair ( ). These groups will form the Morse-Witten Chain groups in the chain complex we are constructing. Denote by ( ) ( } and for { . Observe that for every ( ) { ) } { ( ) the unstable disc , denote by }, the set is finite. We wish to first prove the following: Proposition 4.1 Proof In the case union of ( ) The group , then ( ) is free abelian and finitely generated. ( , so ) ( ). Since then ( ) -discs, all of which are homotopy equivalent to a point, then we see and so is a disjoint ( ) ( ) is free abelian and finitely generated. For the remainder of the proof assume that manifold) we know; . By Theorem 2.8 (Handle decomposition of a ⋃ that is, is obtained from sum of | | -spheres; by attaching finitely many -handles. Therefore ⋁ . Then recall that ( ) ( ) for all is a wedge and we get the series of isomorphisms below: ( ) and so ( ̃( ) ) ) (⋁ ⨁ ̃( ) | ⨁ | ( )is free abelian and finitely generated for all . In the construction of the long exact sequence of the pair ( homomorphism ̂ ( ) ( ) . If natural homomorphism induced from the quotient map ), we have a boundary ( ) ( ( ) ) is the ( ) then their composition ̂ ( ( ) yields a homomorphism chain complex we are constructing. Corollary 4.2 If 0811265 is an Ian Vincent ) ( ) ( ) ( ). This is the desired boundary homomorphism in the -dimensional handlebody, then Page 30 of 50 The inclusion ( ) induces an isomorphism ( ( ) for all ) for all Proof See (Hatcher 2001) Lemma 2.34 parts (b) and (c) on Cellular Homology. Then apply Theorem 3.4 (Handlebodies and Cell Complexes) so there is a 1-1 correspondence between the -handles of and the -cells of the cell complex . We now finish the construction of the Morse-Witten Homology Group: Theorem 4.3 (Morse-Smale-Witten) Suppose that ( ( ) dimensional manifold , and let ( ) and ) is a Morse-Smale pair on a compact ( ) be defined as above. Then: - 1. the sequence ( )→ ( )→ ( )→ → is a chain complex (that is for all ), and 2. for all , there is a natural isomorphism ( ) ( ) where ( ) denotes the th singular homology group of . Proof We follow a similar approach as (Hatcher 2001) does for Cellular Homology. For each of the pairs ( ) ( ) and ( ), we get a long exact sequence. Portions of these long exact sequences fit together into a diagram as shown below: 1. Notice that by the construction above, I have factored and as the compositions ̂ and ̂ respectively. By exactness of the sequence on the diagonal from ‘top-left’ to ‘bottom-right’: ( 0811265 Ian Vincent ) ( ) ̂ ( ) Page 31 of 50 we see that the composition ̂ is zero. Therefore, ̂) ( ̂) (̂ ( ) ̂ ̂ Since the argument works for any , this proves that the sequence ( ) ( ) ( ) ( ) is a chain complex. ( ) for all . 2. We now prove that ( ) with By Corollary 4.2, and using exactness of the diagram above, we can identify ( ) . (̂ ) ( Since onto ( ) ̂) also injective, , this shows that is injective and so it maps ( ) isomorphically onto ) and (̂ ) ( ( ̂ ) isomorphically onto ( ). Therefore ( ( ) ) and ( ( ( ̂ ) isomorphically ( ) ( ̂ ). Since ) and so we have shown ) isomorphically onto is maps . Therefore, it induces an isomorphism ( ( ) ) ( ̂) as required. 4.2 – Geometrical Interpretation The following approach follows from section 3.2 on Morse-Smale Dynamics, and we follow the method outlined in (Salamon, Morse Theory, the Conley Index and Floer Homology 1990): As before, we denote by the gradient-like vector field for – and by ( ) the unstable and stable manifolds for ( ) ( ) { } . ( ) ( )) for every critical point We first choose an orientation of the tangent space ( and denote by 〈 〉 the pair consisting of a critical point and this orientation. For every we then denote by the free group ⨁ 〈 〉 ( ) Claim 4.4 The function being of Morse-Smale type implies that the intersection ( ) ( ) has dimension 1 if the indexes satisfy ( ) . ( ) Proof Transversality of the intersection ( ) ( )) ( ( ) ( ) ( ) since and 0811265 Ian Vincent ( ) ( ) means that the codimensions satisfy: ( )) ( )) ( ( ( ) ( ) we have Page 32 of 50 ( ) ( ( )) ( ) ( )) ( ( ) ( ) ( ) ( ) ( )) As ( ) we have . As the flow lines equal to the intersection ( ( ) ( ), we see that the set of all flow lines between two consecutive critical points has dimension . Proposition 4.5 If the difference of the indices ( ) gradient flow lines of from to . ( ) 1, then there are finitely many Proof Pick local coordinates ( ) for the critical point . In these coordinates, let ( ) centred at with radius . Then we claim that the intersection ( be a ball ( )) is transverse: ( We apply Lemma 2.11 (General Position). Since then there is an isotopy { } of ( ) ( )) by Claim 4.4 and ( ) ( )) such that and ( intersects transversely in finitely many points. Since isotopies will not change the number of flow lines, we see that the set is finite. Each point in corresponds to a gradient flow line of finitely many gradient flow lines of from to . connecting and ; therefore there are ( ) ( ) ) by Since the intersection ( ) is finite if ( ) , define an integer ( assigning a number or to every flow line, and taking the sum. We choose whether to assign +1 or -1 as follows: ( ) Let ( ) be such a flow line (that is ( ) and ( ( )) with Then the orientation 〈 〉 induces an orientation on the orthogonal complement | ( ) ( )| in ( ). In the case ( ) ( ) , the tangent flow ( ) ( ) ( ) ( ) of . induces an isomorphism from ( ) onto ( ) and we define to be or according to whether this ) ∑ map is orientation preserving or orientation reversing. Then define ( where the sum runs over all flow lines connecting to . A boundary operator of the chain complex defined on the previous page is now given by the formula: 〈 〉 ( ∑ )〈 〉 ( ) Of course, we must show that this chain complex gives us the same homology groups as the one as defined in ( ) coincides with defined in section 4.1. We first show that Proposition 4.1 we see that ( ) | can form 〈 〉, the pair consisting of a critical point that ( ( | ( . Now for every critical point and an orientation of ) (the unstable manifold of ( ). By of index , we ( )). Hence we see actually forms a generator of the group ) as we can see in a diagram in local coordinates ( ) as given by the Morse Lemma: 0811265 Ian Vincent Page 33 of 50 ( The bold horizontal line here is a class of boundary is contained within whose . Notice that this is exactly the unstable manifold respect to the usual Morse-Smale pair ( Lemma 4.6 〈 〉 ∑ ), since it is a -cell in ). 〈 〉 with map Consider the Groups defined by ⨁ ( )〈 〉 as above. Then ( ) is a chain complex; that is is isomorphic to the singular homology ( ) with ( ) of the ambient manifold defined by and the quotient . Proof (Sketch) We follow the argument given in the proof of Lemma 3.2 of (Salamon, Morse Theory, the Conley Index and Floer Homology 1990): For simplicity, we assume orientability of . [This is unnecessary for the proof but makes things simpler. (Salamon) gives more information about reducing ourselves to the orientable case. ( ) If is oriented, then for every regular value , the level set is an oriented ( ) will be called positively oriented if submanifold of . More precisely, a basis of ( ) ( ) since ( ) ( ). defines positively oriented basis of ( ) ( ) ( ) ( ) ( ) It follows that the descending sphere inherits an orientation from ( ) ( ) ( ) inherits an orientation from ( ). The integer and the ascending sphere ( ) in ( ) agrees with the intersection number of ( ) and ( ) in ( ). For every critical point orientation of where of , observe that in the previous observation that the choice of an ( ) determines a generator of the relative homology group is the index of . Hence Next, we define ( ) ( ) can be identified with ) ). ( ) ( ) ( ); the union of orbits connecting to . In the case , this is a submanifold of dimension . Define the homomorphism ( 0811265 ( ( Ian Vincent ) ( ) Page 34 of 50 to be the boundary homomorphism as in Theorem 4.3. We need to show that this coincides with the map . Recall that ( ) is the flow of the vector field as before. Denote by , and { ( { ( ( )) ( ) ( )) } ( ) ( ) { } { ( } ( )) } These sets are illustrated the diagram below, consisting of a local picture about the critical points and . In the diagram, the flow lines are shown in red, and respectively are shown in grey, while and are shown respectively in blue: Note that is contractible onto ( ) { } by taking the limit . Likewise, } whose width converges to zero as ( ) { defines a tubular neighbourhood of . { } ( ) ( ) ( ) Since and intersect transversally, it follows that consists of ( ) finitely many components each containing a unique point . That is, there exists a diffeomorphism (its corresponding handle in the handle decomposition of ) ( ) ( ( ) ) { } with and ( ) { } where . In particular | is a -manifold with boundary , and the map ( , diffeomorphic to induces an isomorphism on homology ) ( ) ( ) ( ) determines a generator of the homology The given orientation of ( ) which under the above isomorphism is mapped to a generator homology class is determined by the orientation of via the flow defined isomorphism one inherited from where 0811265 ( ( ). The inherited from the orientation of ( ) ( ) via the injection ( ) ( ) ( ) as the first basis vector). Indeed, both orientations agree if and only if { } is the sign associated to the connecting orbit Ian Vincent ) ( ). This orientation may or may not agree with the ( ) (by taking via ( ) ( ). Page 35 of 50 Now choose a triangulation of the -manifolds and extend it to a triangulation of the } with boundary }. Together with the given orientation of ( ) { ( ) { manifold ( ) this determines a generator ( ) ( ) ( ) ( ) ( ) ( The homology class original triangulation of agrees with ( ) ( ) ) ( ) is represented by the ( ) and therefore together with the orientation inherited from ( . Using the above isomorphism ∑ ( ) ) and therefore proves that the two boundary maps results follow. ( ( and ) we obtain ) coincide, and by Theorem 4.3, the 4.3 – Examples Let us look at some examples. The tricky part of the construction to see whether the isomorphism ( ) ( ) is orientation preserving or orientation reversing. To avoid mistakes I have sketched the local picture of the flow lines at each critical point, and given an orientation locally. Then by performing the flow along each of the flow lines, out of one chart and into another, this enables us to keep track of the orientations: Example 4.7 The Torus Consider a torus embedded in (tilted by an isotopy), with its ) usual Morse function defined by ( . Then there are four critical points with indices respectively. Here is a picture, showing the flow lines from to and in red, and the flow lines of from to in blue. Looking locally at the critical points we see the following local pictures. In the below diagrams we see the red flow lines from to and the green flow lines from to . The blue flow ( ) and ( ). We orient the tangent spaces ( ) ( )) ( ) lines correspond to ( ( ( )) and 0811265 ( ) Ian Vincent ( ( )) as shown on the diagrams by 〈 〉 〈 〉〈 〉 respectively. Page 36 of 50 Now, the orientation 〈 ( ) of | ( ) for ( )| ( ) induces an orientation on the orthogonal complements 〉 on { }. Therefore the orientations are ( ) Let ( ) ( ) be the tangent flow along the flow line 〈 〉 〈 〉 〈 〉 for 〈 〉 ( ) 〈 〉 ( ) 〈 〉 { } then we see 〈 〉 〈 〉 〈 〉 〈 〉 Equivalently, the change of basis matrices are: ( ) Therefore the maps Therefore by definition, ( ) ( are orientation preserving and and , so 〈 〉 ) are orientation reversing. 〈 〉 〈 〉 Similarly, the below diagrams show the local pictures of the flow lines around The orientation 〈 〉 on | ( ) ( )| { for ( ) induces an orientation of the orthogonal complements The orientation 〈 〉 on ( ) ( )| { for ( ) ( ) induces an orientation of the orthogonal complements ( ) be the tangent flow along the flow line 〈 〉 ( ) of } (which consist of single points). Therefore the orientations are ( ) Let ( ) of } (which consist of single points). Therefore the orientations are ( ) | : 〈 〉 〈 〉 for 〈 〉 ( ) { } then we see 〈 〉 〈 〉 〈 〉 〈 〉 Equivalently, the change of basis matrices are: ( Therefore the maps by definition, 0811265 Ian Vincent ) ( are orientation preserving and and , so 〈 〉 ) ( ) are orientation reversing., and so 〈 〉 〈 〉 〈 〉 . Page 37 of 50 It follows that all the boundary maps homology groups are: ( in the chain complex ( ) ( ) are 0 and so the Morse-Witten ) ( ) Which coincide with the singular homology groups of the torus. Example 4.8 The Projective Plane Consider the identification diagram below of with a Morse-Smale Flow induced by the red arrows: and with indices ( ) Observe that there are three critical points ( ) , together and ( ) . Therefore we form a chain complex: 〈 〉 And we now compute the boundary maps 〈 〉 and 〈 〉 : From the above diagram, we get the following local pictures: The red lines denote the flow lines from to , the green lines denote the flow lines from to ( ) and the blue lines denote the flow lines of to . Since ( ) , we do not consider the ( ) ( )) ( ) ( )) and green flow lines. We orient the tangent spaces ( ( ( ) ( )) as shown on the diagrams by 〈 ( Now, the orientation 〈 ( ) of | ( ) for ( )| 〉 on { ( ) induces an orientation on the orthogonal complements }. These orientations are ( ) Also, the orientation 〈 〉 on | ( ) for ( )| 0811265 { 〉 〈 〉 respectively. 〈 〉 ( ) 〈 〉 ( ) induces an orientation on the orthogonal complements }. These the orientations are Ian Vincent Page 38 of 50 ( ) of ( ) Let 〈 〉 be the tangent flow along the flow line 〈 〉 〈 〉 〈 〉 〈 〉 for 〈 ( ) 〈 { } then we see 〈 〉 〈 〉 〉 〉 〈 〉 Equivalently, the change of basis matrices are: ( Therefore the maps Therefore by definition, ) and Morse-Witten homology groups ( ) { } ( ) ( ) are orientation preserving and and , so 〈 〉 ( ) is orientation reversing. 〈 〉 and 〈 〉 〈 〉 so the are: ( ) 〈 〉 〈 〉 ( ) Which coincide with the singular homology groups of the projective plane. 0811265 Ian Vincent Page 39 of 50 ( ) ( ) is, by definition, the set of Example 4.9 The Special Unitary Group Recall that complex matrices satisfying and . By standard techniques in Lie Group ( ) Theory, we see that as a manifold, . To compute its homology, rather than attempting to decompose it into an -dimensional CW complex in the usual method of computing the Homology, we compute the Morse-Witten Homology which given a Morse function, we can form a corresponding Chain complex. ( ) Define by ( ( ) (Matsumoto 2002), page 99, proves that critical points: ). The following Lemma from as defined above is a Morse function, and finds its ( ) is a Morse function, and its critical points Lemma 4.10 The restriction of to consist of the diagonal matrices, whose diagonal consists of s and s, and whose determinant is . Proof The proof is not hard to follow in (Matsumoto 2002), and is out of the scope of the document, so I omit it from here. We compute the indices of the critical points as follows: Writing a critical point of in the form ( Once more, from (Matsumoto 2002), if order), the index of the Morse function If { ), where }. are the subscripts with (in ascending at the critical point is given as follows: ( ) , the index is ( and if ) ( ) ( ) , then the index is ( ) ( ) Therefore, simply by considering the possibilities of the diagonal matrices of this form, we see that the critical points of and their corresponding indices are given below: Critical Point ( As we see, 0811265 ) index ( ) index ( ) index ( ) index has four critical points of indices Ian Vincent Index and . Therefore the chain complex is given by Page 40 of 50 Fortunately, in the above Chain Complex we conclude that all the boundary maps if . Hence the homology groups are: ( ( )) { otherwise Critical Point ( Index ) index ( ) ( ) ( ) ( ) ( ) ( ) ( index index index index index index ) index ( ) is given by: Hence the Chain complex for → → → → → → → As before, we can immediately conclude that all the boundary maps { }, we immediately see Therefore for if ( ( )) { otherwise 0811265 for all ( ) We can apply much the same technique as above with ( ). First note that ( ) . As before, let be the function ). By Lemma 4.10 the critical points of are: Example 4.11 ( ( )) ( ) ( To compute are ( ( )) and Ian Vincent ( are , except for ( )), we would need to work out the boundary map Page 41 of 50 . . Chapter 5: Poincaré Duality While a little off topic, I could not resist showing this proof of Poincaré Duality using Morse Theory, as presented in (Matsumoto 2002). 5.1 – Cohomology Groups We quickly introduce the Cohomology Groups ( ). Consider the chain complex ( )→ ( )→ ( )→ ( ) ( ) → of a cell complex . We may assume using Simplicial or Cellular Homology that the rank of ( ) is equal to the number of -cells contained in a CW-structure for , and that all the groups ( ) are finitely generated. Definition 5.1 (Cochains, Coboundary Homomorphism and Cochain Complex) Let ( ) { ( ) } ( ( ) ) be the set of all homomorphisms from ( ) to ; then ( ) becomes an abelian group with respect to the addition of homomorphisms. Here, the ( ) is defined by ( )( ) ( ) ( ) for all ( ) . The addition for subtraction is defined similarly. ( ) is 1. The group ( ) is called the -dimensional Cochain Group, and each element called a -Cochain. ( ) 2. For any -cochain , the composition ( ) ( ) ( ) is a ( )-cochain. The map with the boundary homomorphism ( ) ( ) where ( ) , is called the Coboundary Homomorphism.. 3. The sequence of the cochain groups and coboundary homomorphisms ← ( )← ( )← ( )← ( )← ← ( )←{ } is called the Cochain Complex of . Lemma 5.2 For any . ( ). Then ( ) ( ) ( But by definition of the chain complex, we have is a homomorphism. Proof Take any ) ( so ) ( ) since ( ) of The kernel is denoted by ( ) and is called the -dimensional cocycle group. This group ( ) is a subgroup of ( ), and the elements of ( ) are called cocycles. Also, the image ( ) is denoted by ( ) and is called the -dimensional coboundary group, which is a subgroup of ( ) as well. The elements of ( ) are called coboundaries. By Lemma 5.2, we have ( ) ( ) ( ). Definition 5.3 (Cohomology Group) Cohomology group and denote it by 0811265 Ian Vincent We call the quotient group ( ) ( ) the -dimensional ( ). Page 42 of 50 ( ) is called a -dimensional cohomology class. The cohomology class to which a An element of -dimensional cocycle belongs to is denoted by [ ]. ( ) Since a -cochain is a homomorphism , for any -cochain and any -chain , an integer ( ) is determined. By the definition of the coboundary homomorphism, we have ( )( ) ( ) From this we can easily show that the value ( ) for a -dimensional cocycle and a dimensional cycle is determined only by the cohomology class [ ] and the homology class [ ]. (In the next section on intersection forms, we will go over a similar argument in some detail, when we prove that an intersection number is determined between homology classes.) Therefore, a ( ) dimensional cohomology class [ ] determines a homomorphism [ ] . We set ( ( ) ) ( ) {homomorphisms } and from the above we obtain a map ( ) ( ( ) ) ( ) to the homomorphism on ( ). This is a homomorphism. The next by assigning [ ] theorem is a special case of a weaker version of the theorem called the Universal Coefficient Theorem. ( ) Theorem 5.4 The homomorphism ( ). the torsion part of ( ( ) ) is surjective, and its kernel ( ) is Proof See the Universal Coefficient Theorem in (Hatcher 2001), pages 194-196. Now we can state the theorem that we will prove in the next section: Theorem 5.5 (Poincaré Duality) For ( ) an orientable closed { } ( ) -manifold , we have 5.2 – How to prove Poincaré Duality First, we make some observations on relations between boundary and coboundary homomorphisms. If a cell complex contains taken as a natural basis of , then the oriented cells 〈 -cells, ( ). A similar basis 〈 ( ) so that the boundary homomorphism these bases. In fact, if (〈 Then 〉) 〈 〉〈 〉 〈 can be represented by a 〉 For the -dimensional cochain group 0811265 is a homomorphism Ian Vincent 〈 〉 〈 〉 can be 〉 can be taken for ( ), ( ) can be represented by a matrix using 〈 〉 integer-valued matrix ( That is, 〉 〉〈 ) ( ), take the dual basis ( ) , of 〈 〉〈 〉 defined by Page 43 of 50 〈 〉. 〉) (〈 { when otherwise ( ) We represent the coboundary homomorphism basis ( ) and the dual basis of ( ( ) as follows: of ( So that ( ) by a matrix using the dual ) ) . ( ) ( ) Lemma 5.6 The matrix which represents the transpose of the matrix which represents ( ) with respect to the dual basis is ( ). 〉. By the Proof To compute the coefficients , we evaluate both sides of the above at 〈 defining equations of the dual basis, the RHS can take a non-zero value only at the th term, and the value is equal to . On the other hand, from the definition of the coboundary homomorphism, the LHS is equal to (〈 and we obtain 〉) ( 〈 〉) ( 〈 〉 〈 〉 〈 〉) . We now delay the Proof of Theorem 5.5 further, and discuss how the boundary homomorphism in Cellular homology can be interpreted in terms of handlebodies. 5.2.1 The Attaching maps of a Handlebody and Intersection numbers Let be an oriented ( closed -manifold. ) We ( ) use a handle ( decomposition ) arranged in increasing order of indices. As we have shown in Theorem 3.4 (Handlebodies and Cell Complexes), can be regarded naturally as an -dimensional cell complex , by regarding the cores of -handles of this handle decomposition as -cells. Choose and fix an arbitrary orientation 〈 〉 for the core of the -handle . The we obtain a chain complex ( )→ ( )→ → associated with the handle decomposition of boundary homomorphism the -cell ( )→ ( ) . Using Cellular Homology for example, the is represented by the degree under the attaching map ( ) of how many times covers of the cell complex, with the orientations taken into account. (Here denotes the -skeleton of more details on this, see (Hatcher 2001) page 140. when regarded as a cell complex.) For When a handlebody is regarded as a cell complex, attaching maps of the cell complex are essentially the attaching maps of the handlebody. Using this fact, the degree can be interpreted in terms of a handle decomposition as follows: In the handle decomposition ( let ) ( ) ( ) be the subhandlebody with all the handles attached up to (and including) 0811265 Ian Vincent -handles. Page 44 of 50 Suppose now that ta ( )-handle is attached to . The image sphere embedded in ( by an attaching map ) of the attaching sphere under . To simplify the notation, we denote this by ( is a - in what follows: ) On the other hand, the belt sphere of a -handle contained in is an ( )-sphere, and also is a submanifold of . To simplify the notation again, denote this belt sphere by The dimension : of the attaching sphere add up to the dimension of and the dimension of the belt sphere . Now, we apply Lemma 2.11 (General Position) to the handlebody { } of the boundary such that and ( , so we can find an isotopy ) intersects the belt sphere transversely at finitely many points. Using Theorem 2.12 (Sliding Handles), the attaching map can be replaced by by this isotopy, so that we can assume, with no loss of generality, that the attaching sphere and the belt sphere intersect transversely. (By applying this argument to the disjoint union of belt spheres, we can assume that any attaching sphere and any belt sphere intersect transversely in .) First we consider a simple situation, where an attaching sphere intersect at a single point in . In this case, as we see by making the -handle thinner to get a corresponding -cell , the attaching sphere or negatively. In other words, we obtain this diagram.] 0811265 Ian Vincent and a belt sphere covers thinner and exactly once, positively . [See the diagram below, noting that Page 45 of 50 in To study general situations, we use the intersection number. The intersection number is defined for two oriented submanifolds 〈 〉 and 〈 〉 intersecting transversely at finitely many points in an oriented manifold 〈 〉. (Here is it assumed that .) The sign of the intersection, or , is assigned to each intersection point of and , and the intersection number between 〈 〉 and 〈 〉, 〈 〉 〈 〉 is the integer obtained by adding up all the signs of all intersection points. We saw this before in the Geometrical interpretation of MorseWitten Homology, where we defined the boundary homomorphism 〈 〉 ∑ ( )〈 〉 Determining the sign of an intersection to be or is a question of orientations. There is a nice way of determining the sign of an intersection using exactness, which we explain now: 5.2.2 Using Exactness to determine the Sign of an Intersection: Suppose and are two oriented submanifolds of diagram below: , and . Then for , consider the where the functions and are all the inclusion maps. In particular, note that this diagram we construct an exact sequence in local coordinates: ( where for that ( ) ( ), ( ) ( ( ) ( )) ) ( ) ( ) ( ( ) ( )) and for ( ( ) ( ) ( ) . From ) , ( ) ( ) ( ) . Note , implying exactness. 〉 for ( ) and another ordered basis 〈 〉 for ( ) Take an ordered basis 〈 ( ) and 〈 〉 is an ordered basis for . Then for each , take [here we identify with its image ( ) since is injective]. Call this basis an induced basis for ) . Declare the sign of this exact sequence to be if, given positive bases for ( ), this induces a positive basis for and ( . Otherwise, the sign of this intersection is . ( ); we know that is surjective (by Note that for , there is no unique choice for exactness) and so at least one preimage exists. Therefore we need to check that the sign is ( ): independent of the choice of preimage ( ) Suppose that for each there exist and with ( ) 〈 possibilities for an induced basis are 〈 〉. We need to show that is a positive basis iff basis; we look at the change of basis matrix between them. 0811265 Ian Vincent . Then two 〉 and is a positive Page 46 of 50 First notice that identifying ( ( ) ( ) ( ) with its image exist so that change of basis matrix from ∑ ) so by exactness, ( ( to ; that is ) (here ). This means that there ∑ and so is of the form ( and hence its determinant is )) ( for each . Therefore the ) is positive iff is positive. Via routine arguments, similar to that above, we can check that the sign is independent of all other ) or another positive basis 〈 〉 choices of equivalent bases; another positive basis 〈 〉 for ( ) will induce a positive-determinant change of basis matrix on the induced bases. for ( Since, in our case, we are given orientations of all the attaching spheres and belt spheres of the handles, applying the method above we can determine (up to sign) the intersection number by giving it an arbitrary orientation. 〈 We note that covers 〉 〈 〉. the same number of times as intersects consideration, in this case. By choosing all the orientations of 〈 〉〈 with signs taken into 〉 and 〈 〉 “naturally”, one can compute that ( Where the ) 〈 〉 〈 〉 make up the matrix representing the boundary homomorphism . The sign ( ) in front is not essential here, so we omit the details, but we explain how to define these “natural” orientations. Choose an arbitrary orientation 〈 〉 for the core of the -handle . If we denote by 〈 〉. Namely, since the cocore of , then an orientation of is naturally determined from 〉 is and intersect transversely at a single point, the orientation 〈 〉 determined in such a way that 〈 〉 〈 under the assumption that an orientation of the whose manifold is specified. In general, an orientation of the boundary of a manifold orientation of . The orientations of attaching spheres the ones naturally induced from . The equality is naturally induced from an and belt spheres ( ) 〈 〉 〈 are chosen to be 〉 holds with all these conventions. 5.3 – The Proof of Poincaré Duality Finally, after all the preliminary discussions above, we are ready to prove Theorem 5.5. Proof of Theorem 5.5 (Poincaré Duality) The key point of this proof, which uses Morse Theory is to turn the manifold “upside down”. For this purpose, we just replace the original Morse function by . As we have already seen in the proof of Theorem 2.16, the set of critical points of the Morse function is identical to the set of critical points of . However, a critical point of index becomes a critical point of index with respect to . This can be seen easily by looking at the 0811265 Ian Vincent Page 47 of 50 standard form of about a critical point. The roles of the core and cocore of a handle are switched, )-handle and a -handle transforms itself into an ( . The core of is , and the cocore is . The handle decomposition associated with is ( ) ( ) ( ) of -handles, -handles,…, -handles, are, respectively the )-handles,…, -handles before it was turned upside down. ( where the numbers numbers of -handles, ( ) Let be the cell decomposition of associated with the handle decomposition ( ). From this cell complex, we obtain a chain complex ( The basis 〈 〉〈 〉 )→ 〈 ( 〉 of ( ) ( )⏞ ( ) ) looks like the dual basis of ( ) when we use intersection numbers. In fact, if we compute the intersection number (in ) between a basis 〉 of ( ) and from the formulae { 〈 〉 〈 〉 element 〈 and by the formula 〈 〉 〈 〉 ( ) 〈 〉 〈 〉}, for any submanifolds For a basis element 〈 〉 of ( ), we obtain 〈 ( ( Furthermore, since , we have} 〉 〈 )) ( 〉 { ( ( )) ( )( ) if if , an isomorphism ) ( ) ) can be defined by assigning ( )( , the dual basis element with sign adjusted to a basis 〉 of ( ). Such an isomorphism element 〈 is defined for each dimension . Lemma 5.7 For any , the following diagram commutes up to sign: Namely, Proof We find a matrix presentation of the boundary homomorphism basis 〈 〉 of ( ) (for all ) and the basis 〈 〉 of with respect to the ( ) (for all ). Set (〈 Let ( (〈 〉) 〈 〉 〈 〉 be the subhandlebody consisting of all handles from -handle through to ) in the handle decomposition ( ). We see that boundary 0811265 〉) Ian Vincent of is equal to the boundary , and the of .Considering the fact Page 48 of 50 is now the attaching sphere of an ( that the form belt sphere is now the belt sphere of an ( , and the former attaching sphere handle )-handle in )- , we obtain ( ) 〈 ( by an argument similar to that of 〉 〈 ) 〈 〉 〉 〈 〉. They look identical, but we need to observe a few technicalities. First of all, the intersection number is counted in 〈 to apply the same argument as when we computed 〉( 〈 〉) in this case. Also, , orientations of and must be defined in such a way that 〈 〉 〈 〉 This is reason behind the symbol in the notations of orientations above. If we use the orientations 〈 sign is ( )( 〉 and 〈 )( ) 〉 based on the “natural” orientations, then the difference in , from the formula 〈 〉 〈 〉 ( ) 〈 〉 〈 〉}. The above considerations imply that if we interchange the order of 〈 compare with the formula for 〉 and 〈 〉 and , then we obtain ( ) In this formula, sign ( ) is not important for the proof of Poincaré Duality. In any case, we have found that the matrix representing is equal to the transpose of the matrix which representing or its negative . ( ) By Lemma 5.6, the matrix representing is the transpose of . Hence it is now clear that ( ) and ( ) by the isomorphism identifying ( ) and ( ) using the isomorphism . ( If we use Lemma 5.7, and consider the isomorphism we find that the chain complex ( ( ) with respect to the dual basis and coincide up to sign by . The same argument applies to )→ ( )→ ( )→ ( )→ ) ( ( ) in each dimension , ) is isomorphic to the cochain complex Therefore, the homology group ( )( ( ) ( )) of dimension of the top chain ( )( ( )) of dimension of the lower complex is isomorphic to the cohomology group cochain complex. [The difference between and in the coboundary homomorphism does not alter the kernel or image, and so does not affect the structures of the cohomology groups.] This proves the Poincaré Duality ( ) ( ). Corollary 5.8 If is an orientable, connected, closed -manifold, then the highest dimensional homology and cohomology groups are both isomorphic to : ( ) ( ) 0811265 Ian Vincent Page 49 of 50 ( ) Proof By Poincaré Duality we have is easily proved from definitions that ( ) ( ( ) ) ( ) Torsion part of ( ) and ( ) ( ). On the other hand, it ( ) and by Theorem 5.4 it follows that as well. { } Giving an orientation of corresponds to specifying a generator of generator is denoted by [ ], and is called a fundamental class of . Remark For any non-orientable connected -manifold ( ) . Such a specified ( ) , it is known that { }. Corollary 5.9 (Duality of Betti Numbers) If is an orientable, connected, closed -manifold, )-dimensional Betti Number coincide: then the -th dimensional Betti number and the ( ( ) ( ) Proof By the Universal Coefficient Theorem, we have ( ( )) ( )). Combining ( with Poincaré Duality, we obtain ( ) ( ( )) ( ( )) ( ( )) ( ) as desired. References Hatcher, Allen. Algebraic Topology. New York: Cambridge University Press, 2001. Kozlowski, Andrzej. Morse Smale Flows on a Tilted Torus - Mathematica Demonstrations. n.d. http://demonstrations.wolfram.com/MorseSmaleFlowsOnATiltedTorus/. Matsumoto, Yukio. An introduction to Morse Theory. American Mathematical Society, 2002. Milnor, J. W. Lectures on the h-cobordism Theorem. Princeton University Press, 1965. Nicolaescu, Liviu. An Invitation to Morse Theory. Springer, 2007. Salamon, Dietmar. “Lectures on Floer Homology.” University of Warwick, 1 December 1997. Salamon, Dietmar. “Morse Theory, the Conley Index and Floer Homology.” Bulletin of the London Mathematical Society, 1990: 113-140. Schwarz, Matthias. Morse Homology. Basel: Birkhäuser Verlag, 1993. Tu, Loring W. An Introduction to Manifolds. New York: Springer, 2011. 0811265 Ian Vincent Page 50 of 50