Analogues of the Atiyah–Wall exact sequences
for cobordism groups of singular maps
Abstract
Classical results of Rohlin, Dold, Wall and Atiyah yield two exact sequences that connect the oriented and unoriented (abstract) cobordism groups and . In this paper we present analogous exact sequences connecting the oriented and unoriented cobordism groups of maps with prescribed singularities. This gives positive answer to a fifteen-year-old question posed by Szűcs and has interesting consequences even in the case of cobordisms of immersions.
Keywords. cobordism; singular maps; exact sequences
Part 0 Introduction
1 Classical results
In the 1950’s much work was done on the determination of the structures and generators of the oriented and unoriented abstract cobordism rings and . One of the main tools for these computations was the existence of two exact sequences which resulted from individual works of Rohlin [rohlin], Dold [dold], Wall [wallcob] and finally from a conceptual method by Atiyah [atiyah].
Recall that the elements of are represented by oriented closed smooth () -manifolds and two such manifolds are cobordant if they together bound an oriented compact -manifold with boundary with matching orientations; the elements of are similar but without orientations. Between these two groups stands the cobordism group of the so-called Wall manifolds (see e.g. [stong]) i.e. closed manifolds whose first Stiefel–Whitney class is the mod reduction of an integer cohomology class; two such manifolds are cobordant if they together bound a compact -manifold with boundary with matching integer first Stiefel–Whitney classes. Now the forgetful homomorphism from to (i.e. the map we get by ignoring the orientation) is the composition of two forgetful homomorphisms.
The classical Atiyah–Wall exact sequences (see [atiyah, theorems 4.2 and 4.3]) are a long exact sequence
(I) |
and a short exact sequence
(II) |
containing these forgetful homomorphisms.
The main result of the present paper is the generalisation of these sequences to the cobordism theory of singular maps which answers an open question of Szűcs proposed in [hosszu, section 19]. The precise statements of this result can be found in theorems 2 and 2 at the end of the next section after recalling and introducing the necessary definitions. Theorem 2 and theorem 2 generalise the sequences (I) and (II) respectively to cobordism groups of singular maps and they will be proved in part I and part II respectively; then in part III we shall apply them to compute various cobordism groups and finally to obtain analogous exact sequences for bordism groups.
2 Cobordism groups of singular maps
Throughout this paper we consider smooth () maps of -manifolds to -manifolds where is a fixed positive integer and is arbitrary. If we want to indicate the dimension of a manifold, we put it in a superindex (i.e. means that is a manifold of dimension ) but in most cases we omit this index. If we do not state otherwise, then we also always make the technical assumption that any smooth map between manifolds is such that and is transverse to the boundary .
Definition 2.1.
Two smooth map germs
are said to be left-right equivalent (in some sources also called -equivalent) if there are diffeomorphism germs and of and respectively such that . The suspension of the germ is the germ
By the singularity class (or simply singularity) of we mean the equivalence class of in the equivalence relation generated by left-right equivalence and suspension. The singularity class of is denoted by .
Observe that in each singularity class the codimension of the germs is fixed but the dimension is not.
Definition 2.2.
For manifolds the product of the diffeomorphism groups of and has an action on the space of smooth maps defined by the formula . Two maps are said to be left-right equivalent if they are in the same orbit of this action, i.e. for some diffeomorphisms of and of . A map is said to be stable if it is in the interior of an orbit, i.e. it has a neighbourhood such that every element of is left-right equivalent to . We define a map germ a stable germ if it is the germ of a stable map.
In the present paper we only consider stable germs. Note that if the germ is stable, then every germ representing the singularity is also stable, hence it is justified to call singularities of stable germs stable singularities.
Remark 2.3.
The restriction to only study maps with stable germs is quite mild, in the so-called nice dimensions stable maps form an open dense subset in the space of smooth maps; see [math].
Let us now fix a set of singularities of stable -codimensional germs.
Definition 2.4.
A smooth map is said to be a -map if all of its germs belong to singularity classes in . For a singularity class a point is said to be an -point if the germ of at is equivalent to ; the set of -points in is denoted by . By a slight abuse of notation we will sometimes write instead of for better readability.
Definition 2.5.
For a singularity class the minimal number for which there is a germ in the singularity class is called the codimension of the singularity and is denoted by .
Remark 2.6.
If is a -map, then is naturally stratified by the submanifolds for all . Here is a submanifold of of codimension .
Example 2.7.
-
(1)
If only contains the class of regular germs (i.e. those with maximal-rank derivative at ), then -maps are just the immersions.
-
(2)
If consists of and all stable singularities of type , that is, singularities with representatives whose derivative at has corank , then -maps are called Morin maps. Further restricting the set of allowed singularities we can obtain e.g. the so-called fold maps (i.e. -maps) and cusp maps (i.e. -maps) where fold () and cusp () are the two simplest types of Morin singularities; see [mor].
In the following we shall work with -maps to a fixed target manifold up to a cobordism relation defined as follows:
Definition 2.8.
We call two -maps and (with closed source manifolds and ) -cobordant if there is
-
(i)
a compact manifold with boundary such that ,
-
(ii)
a -map such that for we have and .
The -cobordism class of is denoted by and the set of all -cobordism classes of -maps to the manifold is denoted by .
The set admits a natural commutative semigroup operation by the disjoint union: if and are -maps, then so is
and the -cobordism class is well-defined. This operation has a null element represented by the empty map, moreover, it is actually a group operation (the inverse of any element in is explicitly constructed in [hosszu]). This way becomes an Abelian group for any manifold .
Next we shall endow -maps with various stable normal structures. These will be defined in the following four points; loosely speaking they are decorations on the stable normal bundle, that is, for a map the stable isomorphism type of the virtual vector bundle . If is such a decoration, then -maps equipped with -structures will be denoted -maps. In this paper we shall work with the following types of -maps:
-
(1)
: Let be a stable group using the definition of Wall [stabgr, section 8.2] which we recall now:
Definition 2.9.
A stable group is defined as the direct limit of a sequence of group homomorphisms where
-
(i)
there are homomorphisms such that the diagram
is commutative where the inclusions in the lower row are natural,
-
(ii)
there is a weakly increasing function tending to such that the map is -connected,
-
(iii)
there are homomorphisms such that the diagrams
and
commute up to conjugation by an element in the component of the identity,
-
(iv)
there is a commutative diagram
where denotes the interchange of factors and is the conjugation by an element whose determinant has sign .
We define -maps to be -maps for which the virtual normal bundle has structure group in the following sense:
Definition 2.10.
For a stable group and a virtual vector bundle we say that has structure group if there is a fixed equivalence class of vector bundles stably isomorphic to such that the structure group of each vector bundle is reduced to and their equivalence is defined as follows. Note that the reduction of the structure group of to also gives the reduction of the structure group of to for any . Here and later on means the trivial rank- vector bundle over any base space. Now two such bundles represetning are said to be equivalent if for some the bundles and are isomorphic as -bundles.
For example if , then -maps are -maps with oriented normal bundles; if , then -maps are just -maps without further conditions. If consists of all possible singularities of -codimensional germs,then any manifold has a -map to uniquely up to -cobordism and its stable normal bundle is just the inverse of the tangent bundle in any -group of ; the same is true for any abstract cobordism of manifolds, hence in this case we have
As another example, if (i.e. -maps are the immersions), then is the cobordism group of immersions of -manifolds into with normal structure group which was first described by Wells [wells]. For various other singularity sets with being mostly or the groups were considered e.g. by Szűcs [analog], [hosszu]111We shall refer many times to [hosszu] which only considers -maps but contains theorems which work with the same proofs for more general stable normal structures as well, hence in this paper we will refer to them in their general form., Rimányi [rsz], Ando [ando], Terpai [2k+2], Kalmár [kal] and Sadykov [sad] (although in some of the papers cited the codimension of the maps is non-positive, unlike in the present paper).
-
(i)
-
(2)
: We define -maps to be -maps with the first Stiefel–Whitney class being an integer class, that is, a -map is a pair where is a -map and is a cohomology class such that its mod reduction is . Following e.g. Stong [stong] we call this type of normal structure a Wall structure and also call -maps Wall -maps. Again, if consists of all possible singularities of -codimensional germs, then any manifold has a -map to uniquely up to -cobordism and its first normal Stiefel–Whitney class is the same as , hence for this we have
-
(3)
for a natural number : -maps are also called -framed -maps and they are defined so that their germs are of the form (for ) and the change of coordinate neighbourhoods always induces the identity on here. This extends the notion of -framed immersions, i.e. immersions equipped with pointwise independent normal vector fields; if , then -maps are just immersions with normal -framing. We can endow -maps with other stable normal structures as well, and this defines -maps. These maps were constructed in [hosszu, definition 9 and remark 10] and a very important property of them is the following: if the target manifold is of the form for a manifold , then we have
(2.1) and the isomorphism is given by a natural correspondence between framed and non-framed maps; see [hosszu, proposition 13].
-
(4)
: Let be a stable group and a vector bundle over the classifying space . Then -maps are generalisations of -maps and we postpone their precise definition to section 6. Intuitively a -map is a -map such that a bundle induced from splits off from its “normal bundle” (the sense in which we mean this will be clarified in section 6); if is trivial of rank , this just gives an -framed -map.
Now if is any of the above four stable normal structures, then the cobordism of two -maps to a manifold can be defined by adding the -structure to the map in definition 2.8 and the cobordism group of -maps to is denoted by .
To form the statements of our main theorems it remains to define a vector bundle which will play the role of in a special type of -maps. We put and denote by the line bundle over induced by the first Stiefel–Whitney class from the tautological line bundle. The notation of this line bundle is due to the fact that if denotes the tautological rank- vector bundle over , then the restriction of over is the determinant bundle . In our second theorem we shall use the rank- bundle .
Now we are in a position to state our main results which are as follows:
Theorem I. For any set of stable singularities and any manifold there is a long exact sequence
where is of the form where is the involution represented by the reflection to a hyperplane, is the forgetful homomorphism and assigns to a cobordism class the class .
Theorem II. For any set of stable singularities and any manifold there is a long exact sequence
where is the forgetful homomorphism and assigns to a cobordism class the class .
Remark 2.11.
Remark 2.12.
Although we made assumptions on the set of allowed singularities which generally exclude the set of all singularities, i.e. bordism groups in general cannot be considered as -cobordism groups, the above theorems still hold for bordism groups as well; see theorem 11.1.
Remark 2.13.
Recall that the (say oriented) abstract cobordism and bordism groups give rise to an extraordinary cohomology theory ; see [atiyah]. Similarly cobordism groups of singular maps also yield cohomology theories: for the singularity sets and stable normal structures used in the present paper we can define extraordinary cohomology functors (see [hosszu, section 19]) and if is a manifold, then we have
for any natural number (cf. the isomorphism (2.1)). We will not use this fact in our proofs, however we note that the exact sequences claimed above are sequences of these cohomology groups, hence they naturally extend infinitely to the right if we put in the place of for negative ’s.
Theorems 2 and 2 will be proved in part I and part II respectively and in both cases we begin by constructing classifying spaces necessary for the proofs, then we define the above long exact sequence of cobordism groups, and finish by showing that the homomophisms in the sequence are the ones we claimed. We shall also see that these exact sequences generalise the classical exact sequences (I) and (II); see remarks 5.6 and 8.4. After this, in part III we will see applications of theorems 2 and 2 to cobordism groups of immersions and Morin maps, and finally, bordism groups.
Note that in the sequence (I) the arrow is the multiplication by , hence for theorem 2 to be completely analogous to it we should have , i.e. , which is indeed so in some special cases but not generally as we shall see. Moreover, is and not although its classical version in (I) is just . This is because embeds into , hence taking the Poincaré dual of the integer first Stiefel–Whitney class of a Wall manifold up to cobordism is independent of which integer representative of we take, however, remark 2.11 implies that is not always mono and so this simplification which works for abstract cobordism groups may not work in the general case.
Theorem 2 is not a perfect analogue of the sequence (II) either: it is actually analogous to [atiyah, theorem 4.3] (if we do not take into account the normal -structure which is meaningless in the case of abstract cobordism groups). Atiyah in his paper then proves that the long exact sequence splits to short exact sequences yielding (II) but, as noted in remark 2.11 above, this is not true generally.
Part I The oriented case
3 Cobordism of -maps
One of the most important ingredients of any type of cobordism theory is an analogue of the Pontryagin–Thom construction which assigns a classifying space to that theory and gives a bijection between the cobordism classes and the homotopy classes of maps to this space. The classifying space of the cobordism groups (where is a set of stable singularities, is a stable normal structure and is an arbitrary manifold) will be denoted by , that is, is the (homotopically unique) space which has the property
where denotes the one-point compactification of and means the (based) homotopy classes of maps from the first space to the second.
If is a stable group, then many explicit constructions of the space exist (see e.g. [rsz], [ando], [hosszu] or [sad]); we will now briefly recall one of them: the so-called strengthened Kazarian conjecture proved by Szűcs in [hosszu]. After this we shall construct the classifying space of -cobordisms.
A stable group is defined as the direct limit of a sequence of groups where has a fixed linear action on (see definition 2.9). We denote by the universal rank- vector bundle with structure group , i.e.
where is the universal principal -bundle.
Definition 3.1.
As before, let be a set of stable singularities and a stable group. Consider the jet bundle over which has as fibre the infinite jet space (i.e. the space of all polynomial maps with constant term). Denote by the subspace of those maps whose singularity at is in and let be the union of the ’s in each fibre of , that is,
where the action of on is one-sided, we do not act on the source space of the polynomial maps. Now the natural map gives us a map . The Kazarian space for -maps is the direct limit
Note that we have a fibration .
Definition 3.2.
The pullback of the vector bundle by the above fibration will be denoted by where is a rank- virtual vector bundle. Since the natural map induces from , we get that the map induces from . The universal virtual normal bundle of -maps is
The key tool to proving our main theorems will be the construction of the classifying space of cobordisms of singular maps using the Thom space of the universal virtual normal bundle (for -cobordisms this is the strengthened Kazarian conjecture [hosszu, corollary 74] and for - and -cobordisms it will follow from our constructions). Now the Thom space of a virtual vector bundle does not exist, but we can still get a well-defined space (up to homotopy) if we apply the infinite loop space of infinite suspension functor of Barratt and Eccles [be] to it. This is a consequence of [hosszu, remark 60 and definition 72] but it is worthwhile to recall its direct construction here.
Definition 3.3.
Let be a virtual vector bundle of rank over a space such that is the direct limit of a sequence of inclusions where for each the restriction can be represented by for a (non-virtual) vector bundle , moreover, the inclusion is a -homotopy equivalence for a weakly increasing sequence tending to (this holds for but also more generally for any virtual bundle over a CW complex). Then we can define as the Thom space , hence also exists. We define the space
Note that here the inclusion induces isomorphism in cohomologies up to the index for each (by the Thom isomorphism), hence the -homotopy type of coincides with that of (by the Whitehead theorem). This means that the -homotopy type of can be defined as that of . Moreover, we also get that the direct limits and are -homotopy equivalent and since tends to this implies that for any the -homotopy type of is that of for some numbers and .
Remark 3.4.
The space is an infinite loop space, hence it naturally defines a spectrum . What is more, the “virtual space” defines an equivalence class of spectra and this is a representative of it; see [hosszu, remark 62].
The strengthened Kazarian conjecture [hosszu, corollary 74] states:
Theorem 3.5.
If is a set of stable singularities and is a stable group, then the classifying space of cobordisms of -maps is homotopy equivalent to .
Remark 3.6.
The cohomology theory mentioned in remark 2.13 is defined for by the spectrum .
Remark 3.7.
Later it will be important for us to have a deeper understanding of the connection between the cobordism group (for a manifold ) and the classifying homotopy classes , so we now give a short description of it based on [hosszu].
If is a -map and is an embedding, then is a so-called -embedding. A -embedding of into a manifold is defined as a triple where is an embedding with normal -structure, is a sequence where the ’s are pointwise independent vector fields along (i.e. sections of the bundle ), is a foliation of dimension on a neighbourhood of which is tangent to along and any point has a neighbourhood on which the composition of with the projection along the leaves of to a small -dimensional transverse slice has at a singularity which belongs to .
Cobordisms of -embeddings of -manifolds to can be defined in the usual way and their cobordism group is denoted by . Now if the number is sufficiently large, then assigning to the -map the -embedding (with vector fields arising from a basis of and foliation composed of the leaves ) yields an isomorphism
This is stated in [hosszu] as theorem 2 and its proof relies on lemma 43 and theorem 1. This is important to note since the proofs of these all only depend on (ambient) isotopies of the embeddings into which leave the stable normal bundles of the maps involved unchanged.
Now forgetting the singularity structure (given by and ) of yields a well-defined cobordism class of an embedding of into with normal -structure. Since these are classified by the Thom space we have the homotopy class of a map corresponding to it such that is the preimage of and its normal bundle is induced from (note that is the one-point compactification of ). Then using the singularity structure we can lift this classifying map through the map induced by the fibration (see definition 3.2) to get a map
where is now the preimage of and its normal bundle is induced from .
The adjoint correspondence identifies with and if is sufficiently large, the latter is identified with since by increasing we only attach large dimensional cells to the target space. Now the homotopy class can be thought of as an element in and assigning to gives an isomorphism
The composition of the two isomorphisms described above show a bijective correspondence between the cobordism classes of the ’s and the homotopy classes of the ’s which forms the isomorphism of theorem 3.5.
Next we shall define the Kazarian space of Wall -maps and describe the classifying space similarly to the theorem above. Consider the first Stiefel–Whitney class (which is a fibration with fibre ) and let be the inclusion of the -cell.
Definition 3.8.
Let be the homotopy pullback of the diagram , i.e. the space we obtain by pulling back the fibration by , and let be the map indicated on the pullback diagram
In other words we define as the subspace in .
This is the direct limit of the spaces that we get in a similar way from instead of . We can pull back the tautological bundle to a vector bundle and this makes the classifying space of rank- vector bundles equipped with Wall structures, that is, those vector bundles for which is the mod reduction of a specific integer cohomology class (this is true since , and corresponds to the mod reduction). Now we can again consider the jet bundle over and take in each fibre the space to obtain a fibration as in definition 3.1.
Definition 3.9.
If is a set of stable singularities, the Kazarian space for -maps is .
We have now a diagram
where all squares are pullback squares. Using this we can define the rank- virtual vector bundle as the pullback of .
Definition 3.10.
The universal virtual normal bundle of -maps is over .
Theorem 3.11.
If is a set of stable singularities, then we have .
Proof. We need an isomorphism for any target . Recall from remark 3.7 that the analogous isomorphism for -cobordisms is obtained by assigning -embeddings to the -maps, then lifting their classifying maps through the map (where is a large number). As we noted in remark 3.7 the isomorphism of the cobordism groups of -maps and -embeddings (see [hosszu, theorem 2]) does not depend on the choice of normal structures, hence it also yields the isomorphism
for sufficiently large, where is defined analogously.
Now if we have an embedding with normal Wall structure, then its classifying map can be lifted to so if it is also a -embedding, then using the pullback square
this inducing map lifts to . So -embeddings can be induced from and since all of the above extnds to maps between manifolds with boundary (in particular to cobordisms), we get a homomorphism from to as in remark 3.7.
To see that this is an isomorphism, note that if we have a map , then its composition with the map to pulls back a -embedding into (by theorem 3.5) which also has a normal Wall structure. Similarly homotopies between maps to pull back cobordisms of -embeddings with normal Wall structures. Thus we obtained the inverse of the above homomorphism , meaning that it is iso and this concludes our proof.
4 A long exact sequence
We shall now connect the classifying spaces of -cobordisms and -cobordisms. Consider the embedding as a fibre over a point in which then factors through (again as the embedding of a fibre):
If we now take the approximations , and of the spaces above and the approximation of , then both squares in the diagram below will be pullback squares:
But then the same is true for the jet bundles over the spaces , and , hence also for the Kazarian spaces and so the suspensions of the approximations of the universal virtual normal bundles also fit into a diagram
with all squares being pullback squares. After these preliminary observations we are ready to state the following (which is the main statement of theorem 2):
Theorem 4.1.
For any set of stable singularities and any manifold there is a long exact sequence
Proof. Let be any contractible subspace and let be its preimage in . If was the embedding of the fibre over a point in , then and the map is a homotopy equivalence . Moreover, if is the restriction of the Kazarian space over , then we have a pullback square
with the vertical arrows being fibrations, hence we have . Similarly if is the restriction of over , then we have a pullback square
which implies that is stably isomorphic to .
Now let be any point and put (which are contractible subspaces). Let the preimages of and under respectively be and (thus we have ), their preimages in be and (thus we have ) and put and (which are now stably isomorphic to ). Then for all we have a cofibraion
(where and are the appropriate restrictions of ). Since the normal bundle of is induced from the normal bundle of it is trivial and so we have , hence the cofibration above has the form
Now applying the functor to the Puppe sequence of this cofibration we get a sequence of maps
And note that here equals . This sequence is infinite to the right by construction, but it is also infinite to the left since the number could be arbitrary and we get the same maps by applying to the ’th suspensions as by applying to the ’st suspensions. We also note that the maps in this sequence commute with the natural maps (for ). Then converging with to infinity yields a sequence
of the direct limits.
If we then fix a manifold and apply the functor to this sequence, then we obtain the long exact sequence of cobordism groups as claimed.
Remark 4.2.
In the proof above we used the cofibrations
for all . Later we will refer to this (and to similar occurences) as a cofibration
Although these are not actual spaces, applying the functor to them yields a well-defined fibration (cf. the proof of [hosszu, theorem 8]) which is a fibration between the corresponding spectra described in remark 3.4.
Remark 4.3.
If we had in theorem 4.1 , then the same long exact sequence could be obtained from the cofibration
by just applying the functor which yields a fibration
and then taking the homotopy long exact sequence of this.
Now to prove theorem 2 we only have to describe the homomorphisms in the above exact sequence.
5 Description of the homomorphisms , and
First fix a manifold , a vector bundle of rank , a closed submanifold of codimension and with normal bundle and put . If we denote by the cobordism semigroup of -manifolds embedded into a fixed manifold with normal bundle induced from (i.e. -embeddings), then we have
Hence the Puppe sequence of the cofibration gives us a sequence of homomorphisms
of cobordism semigroups (which is exact in the stable dimensions).
Lemma 5.1.
Let be a -embedding which yields a decomposition (with and induced from and respectively) and the exponential map can be used to extend to an embedding of the total space of into a small neighbourhood of . Then the boundary homomorphism assigns to the cobordism class the class of which has codimension , is endowed with the natural normal vector field of the sphere bundle and the orthogonal complement of this trivial subbundle in its normal bundle is induced from .
Proof. We obtain in the Puppe sequence by restricting to a tubular neighbourhood of and then factoring by the complement of this restriction, that is, attaching the cone . The next item in the sequence is which we get by gluing the cone to the previously obtained space. In the following we shall always mean the above representations of the spaces and .

If the homotopy class of induces the cobordism class of (such that ), then its image under
is represented by which we get by taking the portion of the image of in the total space of and pushing it toward the special point of the cone , in this way detaching it from , and otherwise leaving the rest of unchanged.

The inducing map is such that the exponential image of the normal bundle (restricted to a small neighbourhood) is the preimage of a tubular neighbourhood of and it is mapped fibrewise bijectively to its image in this neighbourhood. Thus we may assume that the image of coincides with such that and is again a fibrewise bijection from to .
Then we can assume that the image of intersects the base space only in (this intersection is in ) and restricted to the preimage of this intersection coincides with . Hence for any point we have that maps the sphere to and the union of these (that is, ) is the whole . Since is a sphere bundle, its embedding has a natural normal vector field. This trivialises the subbundle of its normal bundle induced from the direction of the suspension in . The orthogonal complement of this in the normal bundle of is then induced from the part of the normal bundle of . This is what we wanted to prove.
If the manifold above has the form and the submanifold is a fibre , then we have , and , hence the sequence above has the form
Corollary 5.2.
The boundary homomorphism in this case is where is the involution which acts on a cobordism class represented by a -framed -embedding such that it inverts its framing vector field and in the orthogonal complement of in (induced from ) it applies the -action of .
Proof. We want to understand the embedding in the lemma above in this special case. Now is a trivial line bundle, so where the are both diffeomorphic to , the maps can be identified with and the normal vector fields giving the -framing of these two maps are opposite since these are the natural (say outward pointing) vectors of a sphere bundle. Thus we may assume that the -framing of is the same as the -framing of and is opposite to it.
What is left is to understand how the orthogonal complement of in the normal bundle is induced from on the two manifolds . If our fixed fibre was for a point , then (with the notation of the previous proof) the tubular neighbourhood is for a small number . Now and the part of induced from is pulled back by the restriction of from the bundle over . Identifying the fibres in in the trivial way then yields that the parts of the normal bundles of are induced by the same map from the bundle over . But then identifying the fibres over (i.e. in ) instead of changes the identification of with by applying the -action in one of them, say in .
Thus we got that both the -framing of and the inducing map of the rest of its normal bundle coincide with those of , on the other hand, the -framing of is opposite to that of and the inducing map of the rest of its normal bundle is composed with the -action in . Hence we have where and for the involution described in our statement above.
Proposition 5.3.
is of the form where is the involution which, when representing cobordism classes by -embeddings (see remark 3.7), composes a map with the reflection to a hypersurface but does not change its orientation.
Proof. Let be the dimension of and put . Then remark 3.7 and theorem 3.11 show that for the group is naturally isomorphic to
where is a sufficiently large number. Observe also that we have and, using the notation of the proof of theorem 4.1, we let be the fibre of over a point in and is its complement in and for we let the virtual bundle be the restriction of over (hence and ). We now think of these Kazarian spaces as finite dimensional approximations of the actual spaces (hence closed manifolds) over which the ’st suspensions of the universal virtual normal bundles exist as (non-virtual) vector bundles, but for simplicity of notation we are not indicating this.
Then the source and target of are
respectively, thus we are in the setting of corollary 5.2 with , , , and with , and so the boundary homomorphism is indeed . The involution here inverts one (say, the last) of the framing vectors of any -embedding into (see remark 3.7) and applies the -action of in the orthogonal complement of this vector in the normal bundle.
But if is large enough, then we may assume for any -embedding to map into a hypesurface with the last framing vector being the normal vector of this hypersurface, hence inverting it corresponds to the reflection. This would change the orientation of the normal bundle but the -action of reverses the orientation in the orthogonal complement too, which means that the orientation of the normal bundle remains unchanged. This finishes the proof.
Proposition 5.4.
is the forgetful homomorphism that assigns to an oriented cobordism class the class of as a Wall map.
Proof. This follows immediately since the map between the classifying spaces is just the inclusion of into (see the proof of theorem 4.1).
Proposition 5.5.
assigns to a cobordism class the oriented cobordism class of restricted to the Poincaré dual of which can be represented by a -codimensional submanifold uniquely up to cobordism, the restriction can be assumed to be a mapping to , it has oriented normal bundle and its cobordism class only depends on the class of , hence this assignment is well-defined.
Proof. The cobordism class is represented by a -map together with a fixed cohomology class such that its mod reduction is . Then we have a commutative diagram
The Poincaré dual of is represented by the codimension- submanifold for a point . Note that the normal bundle of in is trivial and so has a normal vector field which is uniquely defined by a fixed normal vector of in ; we also have , hence is orientable and can be canonically oriented (since the classifying map of the stabilisation of maps to , hence that of maps to the fibre of over ); finally we may assume that intersects each singularity stratum in transversally, thus is a -map.
By (2.1) we have which means that assigning to the cobordism class of the class of gives a map
taking . We now only need to prove that this map is well-defined and is the same as for which it is sufficient to prove that assigns to the classifying map of the classifying map of .
We again refer to remark 3.7: bijectively corresponds to a homotopy class represented by where is a large number and we again consider such a finite dimensional approximation of the Kazarian space over which exists. In the proof of theorem 4.1 we obtained the map in the Puppe sequence inducing by factoring by its restriction over the complement of a tubular neighbourhood of . Recall that is the preimage of (the fibre of over ). Thus we have a diagram
where the top two rows are connected by quotient maps which factor by the restrictions of the Thom spaces over complements of tubular neighbourhoods.
The homomorphism assigns to the homotopy class of the inducing map the homotopy class of its composition with the quotient map . Now pulls back the -embedding (for an embedding ) from the embedding of into and the composition pulls back a representative of from the embedding of into . But note that we have and its -embedding into is . Hence we proved that and this is what we needed.
This finishes the proof of theorem 2.
Remark 5.6.
If the codimension is sufficiently large (compared to ), then we have and for any (non-empty) singularity set . Hence in the case the portion of the long exact sequence in theorem 2 where is sufficiently small (compared to ) looks like
If we fix here , then can be increased arbitrarily and the homomorphisms in this sequence do not change, hence this sequence is infinite both to the right and to the left. Moreover, by the above propositions these homomorphisms can be identified with those in the classical exact sequence (I), hence theorem 2 generalises the sequence (I).
Part II The unoriented case
6 Cobordism of -maps
Fix a stable group which is the direct limit of the groups . Recall that has a fixed linear action on and denotes the universal rank- vector bundle with structure group . Fix also a rank- vector bundle over the space .
Definition 6.1.
Let be a set of stable singularities. We define a -map from a compact manifold to a manifold to be the germ along the zero-section of a -map
where is a rank- vector bundle over and has the following properties:
-
(i)
the differential restricted to any fibre of is injective,
-
(ii)
noting that (i) implies we require that the map inducing the (virtual) -bundle also pulls back from , thus it pulls back from for sufficiently large numbers .
The cobordism of two -maps is defined in the usual way (i.e. as a -map to where both the germ and the map to inducing the vector bundle extend those of the boundary) and the cobordism group of -maps to the manifold is denoted by .
Example 6.2.
In the case and , the above definition is exactly the definition of -framed -maps in [hosszu].
As we saw previously in part I, the main tools for understanding how cobordisms of -maps are induced from the classifying spaces are -embeddings. In order to construct the classifying space of -maps we now introduce the following:
Definition 6.3.
Let be a set of stable singularities. By a -embedding of a manifold into a manifold we mean a quadruple where
-
(i)
is an embedding,
-
(ii)
where and the ’s are pointwise independent vector fields along , i.e. sections of the bundle ; we identify with the trivialised subbundle generated by the ’s,
-
(iii)
is a rank- subbundle of which is pointwise independent of both and and the normal bundle is induced from the bundle over by a map which pulls back from and from ,
-
(iv)
is a foliation of dimension on a neighbourhood of and it is tangent to along ,
-
(v)
any point has a neighbourhood on which the composition of with the projection along the leaves of to a small -dimensional transverse slice has at a singularity which belongs to .
The cobordism of two -embeddings is defined in the usual way and the cobordism group of -embeddings of -manifolds to is denoted by .
Remark 6.4.
Such a -embedding induces a stratification of by the submanifolds
where the are the elements of and denotes the local projection around along the leaves of .
Example 6.5.
If is a vector bundle over , is a -map and is an embedding, then we can define a -embedding of into : We choose an arbitrarily small representative of and put ; the vector fields arise from a basis in ; the bundle can now be viewed as a subbundle of (since for any we can identify with the translate of to ); and is composed of the leaves intersected by a small neighbourhood of (for all ).
Definition 6.6.
-
(1)
The vector fields and the foliation in the above example are called vertical in .
-
(2)
The subsets (for any ) are called horizontal sections.
The rest of this section consists of constructing the classifying space of the cobordism group of -maps in a similar way as sketched in remark 3.7. First we state a lemma and a theorem which connect -maps with -embeddings and give a stabilisation property for these. Their proofs are the direct analogues of the proofs of [hosszu, lemma 43] and [hosszu, theorem 2], we just repeat them here for completeness.
Lemma 6.7.
Let be a set of stable singularities and be a -embedding of into where is a compact manifold and is any manifold. Then there is a diffeotopy of such that is the identity and (the differential of) takes to the vertical vector fields and to the vertical foliation around the image of . The relative version of this claim is also true, that is, if the vector fields and the foliation are already vertical on a neighbourhood of a compact subset , then the diffeotopy is fixed on a neighbourhood of .
Proof. The manifold is finitely stratified by the submanifolds
By the stratified compression theorem ([hosszu, theorem 1], the analogue of the multi-compression theorem of [rs] for stratified manifolds) there is a diffeotopy of which turns the vector fields into vertical vector fields. Therefore we may assume that is already vertical and so we can also assume that the fibres of are horizontal (i.e. they are in the tangent spaces of horizontal sections). Hence we only need to find a diffeotopy that takes the foliation into the vertical foliation and its differential keeps the vector fields vertical.
We will recursively deform into around the images of the strata . First we list some general observations:
-
(1)
If and are such that each of and intersects each horizontal section exactly once, then a bijective correspondence arises by associating the points on the same horizontal section to each other.
-
(2)
If is such that for each subsets are given as in (1), then a family of bijective maps arises. If we have for any two different points , then the union of these bijections gives a continuous bijective map
-
(3)
If the subsets in (2) are submanifolds of such that and are also submanifolds, then the map is smooth. In this case for all points we can join and by a minimal geodesic in the horizontal section , and using these we can extend to an isotopy of (for which and ).
Denote by the orthogonal complement of the bundle in (with respect to some Riemannian metric). Choose a small neighbourhood of in (where denotes the exponential map of ) and for all let and be the intersections of a small neighbourhood of and the leaves of and respectively.
If the neighbourhoods were chosen sufficiently small, then we are in the setting of (3), hence a diffeomorphism arises (with the same notations as above). Note that and are both neighbourhoods of , the map fixes and for all we have . Observe that where the foliations and initially coincide, this method just gives the identity for all . The isotopy we get this way can be extended to a diffeotopy of (by the isotopy extension theorem) and it takes the leaves of to the leaves of around the image of .
Next we repeat the same procedure around , the image of the next stratum, to get a new diffeotopy (that leaves a neighbourhood of unchanged), and so on. In the end we obtain a diffeotopy of which turns into the vertical foliation around the image of and does not change the vertical vector fields .
Theorem 6.8.
For any set of stable singularities and any manifold , if the number is sufficiently large (compared to ), then we have
Proof. Take any number , so any manifold of dimension at most can be embedded into uniquely up to isotopy. We will define two homomorphisms and which will turn out to be each other’s inverses.
I. Construction of .
For a -map we can choose any embedding and define a -embedding as in example 6.5. Define the map to assign to the cobordism class of the cobordism class of . In order to prove that is well-defined, we have to show that the cobordism class of does not depend on the choice of the embedding and the representative of the class .
Claim. If and are two embeddings and the above method assigns to them the -embeddings and respectively, then .
Proof. Because of the dimension condition, and can be connected by an isotopy . We define a -embedding
(again with the horizontal and the vertical vector fields and foliation), which is precisely a cobordism between and .
Claim. If and are cobordant -maps and the above method assigns to them the -embeddings and respectively, then .
Proof. Let be a cobordism between and (where is a vector bundle over a manifold ). Again by the dimension condition, the embedding extends to an embedding . Hence the map is a -embedding of into (the vector fields and foliation are again vertical and is horizontal) and it is easy to see that this is a cobordism between and .
II. Construction of .
If is a -embedding of a manifold into , then by the above lemma we obtain a diffeotopy of that turns and vertical. A diffeotopy of also yields a cobordism of -embeddings, hence we can assume that and were initially vertical. Now we can define to assign to the cobordism class of the cobordism class of the -map for which and for all (where denotes the projection to and now denotes the exponential of the leaf of at ). In order to prove that is well-defined, we have to show that the cobordism class of does not depend on the choice of the representative of the cobordism class .
Claim. If and are cobordant -embeddings of the manifolds and respectively into and the above method assigns to them the -maps and respectively, then .
Proof. We applied a diffeotopy of to turn the vector fields and foliation vertical (for ), this way we obtained the -map . If and are connected by a cobordism , which is a -embedding of a manifold into , then we can apply (the relative version of) the above lemma to obtain a diffeotopy of that extends the given diffeotopies and on the boundary and turns the vector fields and the foliation vertical. Now composing with the final diffeomorphism and the projection to as above, we obtain a -cobordism between and for which and for all .
The constructions of and imply also that they are homomorphisms and by design is the inverse of , hence they are both isomophisms between and .
Now we can describe the classifying space for which we shall need the pullback of by the fibration (see definition 3.1).
Theorem 6.9.
If is a set of stable singularities, then we have .
Proof. We need to prove that cobordisms of -maps to an arbitrary fixed manifold bijectively correspond to homotopy classes of maps . Since we have and for sufficiently large numbers , it is enough to prove
if is large enough. As in the proof of the previous theorem we will define two homomorphisms and which are each other’s inverses as we shall see.
I. Construction of .
Let be a -embedding of a manifold into . The normal bundle of is of the form and by definition it is induced from by a map . The embedding can be written as the composition
where the disk bundles and are viewed as closed embedded tubular neighbourhoods with and the fibres of are identified with the intersections of the leaves of with the tubular neighbourhood .
Now the vector fields can be extended to by translation along the fibres and if is the leaf of at a point , then we can use its exponential to define a foliation of dimension on which is tangent to the extended along the disk . Applying this for all points we obtain the structure of a -embedding on . Hence there are inducing maps and that pull back from the embedding (again we are considering such a large number and such a finite dimensional approximation of the Kazarian space that the space exists).
The normal bundle of in the disk bundle is induced by the above and this all fits into a diagram
Since induces the normal bundle of , we have in the above diagram , that is, the inducing map factors through the Kazarian space . But then we have a diagram
where all squares are pullback squares.
Considering now the sum , we get , hence the embedding of into is induced by a diagram
This way we can assign to the -embedding the map and since we can apply this construction to cobordisms as well, the assignment is well-defined.
II. Construction of .
Suppose we have a map and put and . Now the normal bundle of in is of the form where and are induced from and respectively by the map .
If we take the preimage of the disk bundle in (of a sufficiently small radius), then we get an embedding where is the normal bundle of and the disk bundles are again considered as closed tubular neighbourhoods. Now we can project to the orthogonal complement of in it along the fibres of and if we denote this projection by , then the composition is a map of to such that the preimage of is , hence it gives the structure of a -embedding to .
We obtain vector fields along by restricting the vector fields along . We also have a foliation of dimension on a neighbourhood of that is tangent to along and the projection maps each leaf of again into a leaf of because of the definition of the inducing map of the -embedding . So we can define a foliation of dimension by defining the leaf of at a point to be the preimage of the leaf of at under .
This yields the structure of a -embedding on the embedding of into . The same method assigns to a homotopy of a cobordism of , hence we can define .
The above constructions imply that and are homomorphisms and also that they are inverses of each other, hence we have proved and this is what we wanted.
7 Another long exact sequence
We shall apply the results of the section above in the case and (recall that was the line bundle over which induces over for all ). The pullback of over the Kazarian space is then and so the classifying space of -cobordisms is .
Remark 7.1.
Note that informally a -map can be thought of as a “-map” with normal bundle of the form . Now this coincides with hence is a map with such a virtual normal bundle from which two (non-virtual) line bundles isomorphic to its determinant bundle split off.
The rest of part II is mainly devoted to the proof of theorem 2. This will be quite similar to the proof of theorem 2 in part I.
Theorem 7.2.
For any set of stable singularities and any manifold there is a long exact sequence
Proof. Recall the pullback diagram
from definition 3.8. The complement of in deformation retracts to which is also a deformation retract of , hence by pulling back its embedding by the fibration to the embedding of a space we get a homotopy equivalence as shown on the diagram
Next we pull back the bundle by the homotopy equivalence to a bundle , thus we have . Then the pullback of the universal virtual normal bundle over , which will be denoted , is stably isomorphic to .
Take the virtual bundle over (which is the restriction of ) and the cofibration
for any . The normal bundle of is induced from the normal bundle of by the composition of with the fibration . The normal bundle of is , this induces over and this finally induces over the Kazarian space , hence the cofibration above has the form
Now applying the functor to the Puppe sequence of this cofibration we get a sequence of maps
This sequence is infinite to the right by construction, but it is also infinite to the left since the number could be arbitrary and we get the same maps by applying to the ’th suspensions as by applying to the ’st suspensions. We can now converge with to infinity since the maps in this sequence commute with the natural maps induced by the inclusions (for ) and so we get a sequence of maps
If we then fix a manifold and apply the functor to this sequence, then we obtain the long exact sequence of cobordism groups as claimed.
Remark 7.3.
Similarly to the oriented case, if we had , then the same long exact sequence could be obtained by turning the cofibration
(cf. remark 4.2) into the fibration
by the functor and taking its homotopy long exact sequence.
Observe that the restriction of over the -cell is trivial, hence the pullback of over will also be trivial which then induces the trivial bundle over the Kazarian space as well. Thus analogously to the proof of theorem 7.2 we get a cofibration
and iterating this process yields:
Corollary 7.4.
For any set of stable singularities and any manifold and integer there is a long exact sequence
Remark 7.5.
It would be tempting to try finding homomorphisms which complete the commutative diagram
with the dashed arrows for different numbers and , however, it seems that such homomorphisms do not exist in general.
8 Description of the homomorphisms , and
Proposition 8.1.
is the forgetful homomorphism that assigns to the cobordism class of a Wall map the unoriented cobordism class of .
Proof. This follows immediately since the map between the classifying spaces is just the inclusion of into (see the proof of theorem 7.2).
Proposition 8.2.
assigns to a cobordism class the cobordism class of restricted to the Poincaré dual of which can be represented by a -codimensional submanifold uniquely up to cobordism, the restriction has a normal -structure and its cobordism class only depends on the class of , hence this assignment is well-defined.
Proof. This is completely analogous to proposition 5.5. For an unoriented cobordism class represented by a -map the Poincaré dual of
is represented by the preimage . Then choosing two distinct hyperplanes and in the Poincaré dual of is represented by
Now the same argument as in the proof of proposition 5.5 yields that assigns to the cobordism class of the class of .
This finishes the proof of theorem 2.
Remark 8.3.
Although there are no apparent interesting algebraic properties of the homomorphism , it has a nice geometric description. Let be the dimension of and put . Observe that we have
and
where is sufficiently large, moreover, is obtained as the boundary homomorphism in the Puppe sequence of classifying spaces. Hence we can apply lemma 5.1 by setting the , , , and in the lemma to be , , , and respectively. This yields that for any cobordism class its image is represented by the mapping of (that is, the circle bundle of the -part of the normal bundle of ) by the restriction of a representative of , together with its natural outward normal vector field.
Remark 8.4.
As in remark 5.6 consider again the case where the codimension is large (compared to ). Then for any (non-empty) singularity set we have and , moreover, we also have since for any embedding with normal bundle this coincides with which is isomorphic to , so this normal structure only depends on and if is large enough, can be embedded into uniquely up to isotopy.
Thus theorem 2 gives an exact sequence for which the portion where is sufficiently small (compared to ) looks like
and if we increase with a fixed , then the homomorphisms do not change which means that this sequence is infinite both to the right and to the left. Then the propositions above show that this sequence can be identified with that in [atiyah, theorem 4.3]. Since the codimension was assumed to be large enough, the manifold constructed in [atiyah, theorem 4.4] can also be mapped to uniquely up to isotopy which now gives a splitting of this long exact sequence yielding the classical exact sequence (II), hence theorem 2 really generalises the sequence (II). Later, in proposition 9.3 we shall show that such a splitting cannot exist in general.
Part III Consequences and applications
In the following two sections we shall apply our two exact sequences for the simplest types of singularity sets, i.e. for (the case of immersions) and , and so on (the case of Morin maps). Our first goal will always be to describe how the endomorphism in theorem 2 acts rationally. By [hosszu, proposition 90] we have
in particular , hence in order to understand rationally we only have to know how the involution which induces acts on the Kazarian space .
After this rational description we shall compute in both cases some cobordism groups of Wall -maps for the above singularity sets which is interesting because Wall cobordism groups are what connect oriented and unoriented cobordism groups. When viewing manifolds abstractly, the classical sequences (I) and (II) yield an obstruction for an unoriented cobordism class can be representable by an orientable manifold, namely that it should be a Wall cobordism class. Now the forgetful homomorphism in theorem 2 is not always injective but otherwise we have the same property, namely for an unoriented -cobordism class to be representable by an orientable -map it should be the -image of a Wall -cobordism class.
9 Immersions
In this section we investigate the case , i.e. -maps are the -codimensional immersions. We shall use the notation
for any stable normal normal structure and any manifold (except that for we decrease and increase by the rank of ). Now the oriented Kazarian space is and the universal normal bundle over it is and we immediately obtain:
Proposition 9.1.
The endomorphism of is the following:
-
(1)
if is even and is the number of non-negative integers such that is odd and , then is trivial on generators and the multiplication by on the rest of the generators (in an appropriate basis),
-
(2)
if is odd, then is the multiplication by .
Proof. We have
where the are the Pontryagin classes and is the Euler class of . The identification of with the degree- part of this graded ring follows from the stable Hurewicz homomorphism (which is rationally iso), the universal coefficient theorem and the Thom isomorphism corresponding to the bundle .
By proposition 5.3 (see also corollary 5.2) the involution inducing acts on this vector bundle so that it inverts the summand and reverses orientation on the summand , thus the Pontryagin classes are unchanged by its action and the Euler class is mapped to . Our claim easily follows from this.
We can say more about these cobordism groups in those cases where the codimension is either small or relatively large.
Small codimensional immersions
First, if the codimension is , then the normal bundle of any immersion is induced by from . The classifying spaces of , , and are , , and respectively and since we have (for all and ) these classifying spaces are , , and respectively.
Remark 9.2.
The group is the cobordism group of those immersions whose normal bundle splits to the sum of three identical line bundles, i.e. the group of -codimensional skew-framed immersions which arise naturally in the study of framed immersions; see [ae].
Now if we view immersions to Eucledian spaces, then theorems 2 and 2 yield the long exact sequences
and
of stable homotopy groups (which, of course, could be both obtained in much easier ways too). This means that in this case our main theorems are not new, however it also shows the following important property of our two exact sequences:
Proposition 9.3.
This is interesting since the classical analogue of is always zero yielding that the long exact sequence splits to short exact sequences of the form (II). However, this proposition shows that the general sequence in theorem 2 does not split to short exact sequences.
Proof. Consider the case of -codimensional immersions of -manifolds. Then the combination of the above two exact sequences gives a diagram
with exact row and column. Since and are trivial (see [liu]) we get that is epi and is mono and since is (see [wu]) this means that is non-zero.
Now let us consider immersions of codimension . In this case the classifying space of is and since coincides with the tautological complex line bundle over this space is . The first few stable homotopy groups of were computed by Liulevicius [liu] and Mosher [mosh] and are as follows:
The action of the involution in theorem 2 for now immediately follows for from proposition 9.1, hence we have:
Proposition 9.4.
The endomorphism of is
-
(1)
on the free part for all and on the free part for all ,
-
(2)
on the torsion part for .
Corollary 9.5.
The cobordism groups for are
where denotes the existence of a short exact sequence .
Large codimensional immersions
Let us now turn to cobordisms of -codimensional immersions such that the dimension of the source manifolds is not much greater than . The reason for this is that for close to the group is close to the abstract cobordism group of -manifolds with normal -structures; see the works of Koschorke [kos], Olk [olk], Pastor [pas] and Li [li] where these groups were computed for and .
Remark 9.6.
For we have and since a cobordism between two -manifolds is at most a -manifold which, when generically mapped with codimension , is immersed.
We shall use the natural forgetful homomorphisms
that assign to the cobordism class of an immersion (or a germ ) with the appropriate normal structure the abstract cobordism class of . These homomorphisms commute with the exact sequences in theorems 2 and 2 and the classical exact sequences (I) and (II), that is, we have commutative diagrams
and
where the rows are exact.
Lemma 9.7.
The map is an isomorphism for all .
Proof. We define the inverse of in the following way: for a cobordism class represented by take the bundle and map it generically to ; let the germ of this map along be and assign the cobordism class to . Such a correspondence would be the inverse of if it was well-defined since is also the determinant bundle of the stabilisation of . So we only have to prove that is an immersion germ and its cobordism class only depends on that of .
Let be another representative of and let be a compact manifold with boundary such that . Then the restriction of to the boundary gives and and if and are their generic map germs to as defined above, then we can extend their representatives and by a generic map
By the dimension condition such a map is stable. Its singular locus is a submanifold of and a straightforward computation with jet bundles shows that we have . Now since the dimension of is at most we generically have , that is, the map in a small neighbourhood of the zero-section is an immersion. This means that the germ of along satisfies both conditions in definition 6.1, hence both and are -immersions and is an immersed -cobordism between them. This finishes the proof.
Proposition 9.8.
For all we have
-
(1)
if is even,
-
(2)
if is odd.
Proof. By Koschorke [kos, theorem 10.8] and Pastor [pas, theorem 3.1] we have
where denotes if is even and if is odd. Now by proposition 9.1 and lemma 9.7 the two diagrams above lemma 9.7 take the form
and
where is the -torsion subgroup of . The -lemma (or the snake lemma) applied to the first of these diagrams implies that is epi and its kernel is and the second diagram implies that this kernel is also a direct summand.
Remark 9.9.
Proposition 9.10.
For all we have
-
(1)
if ,
-
(2)
if ,
-
(3)
if ,
-
(4)
if .
Proof. The first thing to note (similarly to the proof above) is that by lemma 9.7 and the snake lemma applied to the second diagram above it we have an epimorphism . Moreover, if is not a power of and , then by [kos, theorem 10.8] is a direct summand in and so if the image of is non-zero in , then is independent in of the subgroup generated by the -preimage of .
In the following we shall always use the first diagram above lemma 9.7 which yields (again by the -lemma) that is epi in all cases. We will also use that by proposition 9.1 the endomorphism is on the direct complement of in and we obtain the forms of and from [li, theorem 6] and [kos, theorem 10.8] respectively.
Proof of (1). If , we have
hence is either or . But maps onto , so it can only be which can only be a direct summand.
Proof of (2). If and is not a power of , then we have and by [li, p. 472] the involution in theorem 2 is the identity, hence the cokernel of is . If is a power of , then we have and so in both cases we get the diagram
which implies that we have . The part cannot be anything else than a direct summand in but the part is also a direct summand since by the proof of [pas, theorem 3.2] it is generated by a cobordism class which maps to the generator of the part of under the homomorphism induced by the inclusion and this is non-zero in as well, hence the class of in cannot vanish either.
Proof of (4). If , we have
and so is either or depending on whether the in theorem 2 swaps the two summands in or not. Now again the proof of [pas, theorem 3.2] shows that one of these summands is generated by a cobordism class which maps to the generator of the part of under the homomorphism induced by the inclusion . Since the reflection to a hyperplane commutes with this inclusion we get which means that does not swap the summands.
Hence we have and we can also see that is a direct summand in . Now if one of the ’s was not a direct summand in , that would mean that its generator coincided with where is in the preimage under of a non-zero element in , i.e. is an immersion where is not zero. This would imply that all elements in the preimage of were of order or , hence to see that both ’s are direct summands it is sufficient to prove that the preimage under of any element in contains an element of order .
By [wallcob, section 4] we have that the -algebra is generated by cobordism classes , and (for ) of dimensions , and respectively, moreover, and are orientable while is not cobordant to orientable manifolds. The monomials of rank formed by these generators are a basis of the -vector space . We also have and in the above basis the summands and are generated respectively by the monomials containing an even and an odd number of ’s since contains all classes representable by orientable manifolds.
Now is independent of the ’s in which means that all basis elements of which are representable by orientable manifolds have elements of order in their -preimage. Thus it is enough to prove that this also holds for the unorientable basis elements, that is, any monomial of rank formed by the manifolds , and such that it contains an odd number of ’s can be immersed into representing a cobordism class of order .
Observe that is even dimensional for all and is odd, hence such a monomial can be written in the form with and . We can assume that is even and then (1) and (3) yield that there is an immersion such that is in . But then choosing any immersion we get an immersion
which represents an element of order in since we have and is a bi-graded ring. This finishes our proof.
10 Morin maps
Recall that a stable map induces a stratification of according to its Thom–Boardman types (see [boar]): denotes the submanifold consisting of the points where the rank of derivative of drops by , then we have and so on; this defines for any decreasing sequence . In the present section we shall consider maps which only have singularities of type , i.e. Morin singularities (besides regular germs) and we call such maps Morin maps.
Although generally the Thom–Boardman stratification is coarser than the singularity stratification, this is not the case for Morin maps, since a Morin singularity can only be of type (where the number of ’s is ) for some and each type contains precisely one singularity class; see [mor]. We shall denote the singularity class in also by the symbol and note that for we have , that is, Morin singularities form a single increasing sequence. We call Morin maps with at most singularity -maps and put
for any stable normal structure (except that for we decrease and increase by the rank of ). Here is also allowed and it means that we put no further restriction on Morin maps.
The rational cohomology of the Kazarian space of oriented -maps (for ) was computed by Szűcs [hosszu] and we can use this to obtain the form of rationally:
Proposition 10.1.
The endomorphism of is the following:
-
(1)
if is even, is also even and is the number of non-negative integers such that is odd and , then is trivial on generators and the multiplication by on the rest of the generators (in an appropriate basis),
-
(2)
otherwise is the multiplication by .
Proof. Put with . By [hosszu, theorems 6 and 7] we have
with and . Now following the action of the involution inducing the of proposition 5.3 through the spectral sequences computed in [hosszu] yields that changes the and here analogously to how it did in the proof of proposition 9.1, that is, we get and . This implies our claim.
As in the case of immersions, we only have an understanding of the torsion parts of these cobordism groups if the codimension is either small or relatively large. In particular we will investigate the cases when either we have and or is large compared to and is .
1-codimensional Morin maps
Morin maps of codimension were considered by Szűcs [nulladik] who computed for modulo small torsion groups. We also note that is finite -primary for all and by [szszt, theorem 1]. In the following we denote by the Serre class of finite -primary Abelian groups.
Proposition 10.2.
The endomorphism of is the multiplication by modulo (that is, on the odd torsion and free parts) for .
Proof. We proceed by induction on . Observe that setting we get the case of -codimensional immersions where we saw this claim to be true in the previous section. Now assume it to be true for and prove for .
We use the key fibration [hosszu, section 16] to obtain an exact sequence
where the first arrow is the natural forgetful homomorphism and the second one assigns to a cobordism class the cobordism class of its restriction to the most complicated stratum which is an immersion with normal bundle induced from (the universal normal bundle of the -stratum of oriented maps). These both commute with the reflection to a hyperplane (without changing orientation), i.e. the involution in theorem 2.
By the considerations in [nulladik, section 2.2] for odd we have and for even the forgetful homomorphism
(which assigns to the cobordism class of a framed immersion its class as an immersion with normal bundle induced from ) is a -isomorphism. Observe that this homomorphism also commutes with the action of .
In the previous section we saw that theorem 2 for -codimensional immersions is just the long exact sequence induced by the cofibration (where denotes the degree- map), hence in that case the endomorphism of is which means that acts identically on . By the induction hypothesis is the identity modulo on too, thus it acts identically also on modulo . But then the exact sequence above shows that has to be the identity modulo on as well which means that is modulo .
Remark 10.3.
Using [nulladik, theorem A], a similar argument as in the proof above also shows that for we have not only modulo but even on the -primary part.
Corollary 10.4.
For all and the group is finite -primary.
Proof. This is immediate from the proposition above since is an isomorphism on the odd torsion part and the kernel and cokernel of on the free part are and respectively.
This describes generally. Let us now consider the case , i.e. the so-called fold maps and put . Restricting the exact sequence of theorem 2 for fold cobordisms to the -primary parts [nulladik, theorem A] yields
where denotes the Kahn–Priddy homomorphism which maps onto the -primary torsion part of (see [kp]).
Recalling the first few stable homotopy groups of computed by Liulevicius [liu] we obtain
using the notation with summands. This implies the following:
Proposition 10.5.
The cobordism groups for are
where again denotes the existence of a short exact sequence .
Large codimensional fold maps
In the following we shall consider fold maps of codimension such that the dimension of the source manifolds is or and, as we also did above, use the notation
for any stable normal structure (except that for we decrease and increase by the rank of ). The cobordism groups and were determined in these dimensions by Ekholm, Szűcs, Terpai [eszt] and Terpai [2k+2]222As Terpai recently noted [2k+2, theorem 4.a)] is false. We shall elaborate more on this in remark 10.11..
Remark 10.6.
The analogue of remark 9.6 for fold maps is true for : if , then generic maps of -manifolds and -manifolds with codimension are stable and a computation with jet bundles implies that for any such map the codimension of the -stratum and the -stratum in the source manifold are and respectively, hence for a generic -codimensional map of a cobordism of -manifolds is fold. Moreover, and are precisely the dimensions where fold cobordisms are not generic (that is, is a priori not the same as the abstract cobordism group) but they are not very far from being generic as the only other type of singularity that can generically occur is the cusp, i.e. .
As for the case of immersions, we will use the natural forgetful homomorphisms
that assign to the cobordism class of a fold map (or a germ ) with the appropriate normal structure the abstract cobordism class of . These again commute with the exact sequences in theorems 2 and 2 and the classical exact sequences (I) and (II), that is, we have commutative diagrams
and
where the rows are exact.
Lemma 10.7.
The map is an isomorphism for all .
Proof. This is almost completely analogous to lemma 9.7 by changing immersions to fold maps and to . The only thing we have to add to its proof now is that the generic map
(where is a cobordism of -manifolds and ) only has fold singularities in a neighbourhood of the zero-section and its germ along satisfies the condition in definition 6.1 that the differential restricted to any fibre of is injective.
We may assume that the singularity strata intersect transversally which firstly means that is disjoint from the cusp stratum (since the cusp stratum is at most -dimensional and is -codimensional in ), hence it is indeed a fold map on a small neighbourhood of . Secondly it means that the local trivialisations of in a sufficiently small neighbourhood of the zero-section can be chosen such that for all the whole fibre (in this neighbourhood) of containing belongs to , that is, both the fibres of and the fold stratum are orthogonal to .
Now for any point a coordinate neighbourhood of has the form where and are coordinate neighbourhoods in and in respectively. By [rsz, theorem 6] the normal bundle of is induced from over with . The fibres of this in the coordinate neighbourhood of can be assumed to be tangent to for all and can also be identified over all points . Thus in this neighbourhood the map has the desired form described in definition 6.1. If was instead in then is an immersion near , hence then we also get the desired form of and so the first condition in definition 6.1 is always satisfied.
Proposition 10.8.
For all we have
-
(1)
if ,
-
(2)
.
Proof. We shall use the first diagram above lemma 10.7 and obtain from [eszt, theorem 1].
Proof of (1). If is even, then is an isomorphism in this dimension yielding the diagram
and now the -lemma implies the isomorphism claimed.
For odd we have (with an appropriate number ) and this gives
thus is for some depending on how acts on the part of . Now applying lemma 10.7, part of the second diagram above it has the form
hence we have which is isomorphic to a factor group of . But is -primary so this can only happen if , that is, if is an isomorphism.
Proof of (2). We have and and their endomorphisms and are both the multiplication by , hence we get the diagram
which proves our statement.
Proposition 10.9.
For all we have
-
(1)
is isomorphic to a subgroup of of index if ,
-
(2)
.
Proof. We will again always use the first diagram above lemma 10.7 and we know and from [eszt, theorem 1] and [2k+2, theorem 4] respectively.
Proof of (1). If is even, then is an isomorphism and the embedding of into as an index- subgroup.
If is odd, then we have but, as we saw in the proof of proposition 10.8, the part is mapped isomorphically by the endomorphism , hence it does not appear in its kernel. Moreover, for odd embeds into as the kernel of the homomorphism which maps a cobordism class to the normal Pontryagin number . Now if we take the cokernel of , that is, we factor by and by , then the image of under this quotient map will again become an index- subgroup.
Hence both for odd and for even we obtain that there is a group such that the commutative diagram
holds where the monomorphism from into has cokernel . This implies our statement.
Corollary 10.10.
For all the homomorphism embeds into as a subgroup of index .
Proof. We apply lemma 10.7, the second diagram above it and also that the homomorphism is zero as we saw in the proof of proposition 10.8. Then we have
thus the proposition above implies that is a monomorphism with cokernel .
Remark 10.11.
This corollary contradicts [2k+2, theorem 4.a)] which essentially claims that both the kernel and the cokernel of
are . The problem with this, according to Terpai, is the following: the computations in [2k+2] show that here appears as the factor by of the group of a complete set of invariants of cusp null-cobordisms of fold maps (recall that a null-cobordism is a map of a -manifold with codimension and such a map generically only has (fold) and (cusp) singularities); one of the summands (which gets cancelled by the factoring) measures the number of components of the cusp curve of such a null-cobordism modulo and the other one measures whether the orientation changes over the cusp curve or not. However, if we apply [2k+2, lemma 2] correctly, it yields that this second does not appear as the orientation cannot change over this curve. Hence is actually the factor of only by , that is, is mono. Moreover, it is also apparent that its image is the kernel of the Thom polynomial of cusp singularity which maps a cobordism class to the normal Stiefel–Whitney number .
11 Bordisms
In this final section we point out that theorems 2 and 2 also hold when we change -cobordism groups to general bordism groups (although the set of all possible singularities of -codimensional germs generally contains unstable ones which we excluded from ). This can most easily be seen by using somewhat more refined forms of the same arguments as in remarks 5.6 and 8.4. This way we get the following:
Theorem 11.1.
For any CW complex there is a long exact sequence
and a short exact sequence
where is the multiplication by , and are the forgetful homomorphisms, assigns to the bordism class of the map of a Wall manifold the bordism class of and assigns to the bordism class of a map the bordism class of .
Proof. The ’th bordism groups of any CW complex are the same as those of its -skeleton, hence for the purpose of this proof we may only take a finite dimensional skeleton of instead of the actual . A finite dimensional CW complex is homotopy equivalent to an orientable manifold (by embedding it into a Eucledian space and taking a small neighbourhood of it there), and so we may also assume that is an orientable manifold of dimension .
Fix a number and let be such that generic maps of -manifolds to -manifolds are embeddings. Then we have
for any set of stable singularities of codimension .
Claim. .
Proof. Let be any map of an -manifold to and put . Now let be any map that extends the map given on the zero-section of (such an extension is unique up to homotopies fixed on the zero-section) and let be any embedding (which is also unique up to isotopy). Then assigning to t