This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Analogues of the Atiyah–Wall exact sequences
for cobordism groups of singular maps

András Csépai ELTE Eötvös Loránd University, Budapest, Hungary
Abstract

Classical results of Rohlin, Dold, Wall and Atiyah yield two exact sequences that connect the oriented and unoriented (abstract) cobordism groups Ωn\Omega_{n} and 𝔑n\mathfrak{N}_{n}. 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 Ω\Omega_{*} and 𝔑\mathfrak{N}_{*}. 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 Ωn\Omega_{n} are represented by oriented closed smooth (CC^{\infty}) nn-manifolds and two such manifolds are cobordant if they together bound an oriented compact (n+1)(n+1)-manifold with boundary with matching orientations; the elements of 𝔑n\mathfrak{N}_{n} are similar but without orientations. Between these two groups stands the cobordism group 𝔚n\mathfrak{W}_{n} of the so-called Wall manifolds (see e.g. [stong]) i.e. closed manifolds whose first Stiefel–Whitney class is the mod 22 reduction of an integer cohomology class; two such manifolds are cobordant if they together bound a compact (n+1)(n+1)-manifold with boundary with matching integer first Stiefel–Whitney classes. Now the forgetful homomorphism from Ωn\Omega_{n} to 𝔑n\mathfrak{N}_{n} (i.e. the map we get by ignoring the orientation) is the composition Ωn𝔚n𝔑n\Omega_{n}\to\mathfrak{W}_{n}\to\mathfrak{N}_{n} of two forgetful homomorphisms.

The classical Atiyah–Wall exact sequences (see [atiyah, theorems 4.2 and 4.3]) are a long exact sequence

ΩnΩn𝔚nΩn1\ldots\to\Omega_{n}\to\Omega_{n}\to\mathfrak{W}_{n}\to\Omega_{n-1}\to\ldots (I)

and a short exact sequence

0𝔚n𝔑n𝔑n200\to\mathfrak{W}_{n}\to\mathfrak{N}_{n}\to\mathfrak{N}_{n-2}\to 0 (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 (CC^{\infty}) maps of nn-manifolds to (n+k)(n+k)-manifolds where kk is a fixed positive integer and nn is arbitrary. If we want to indicate the dimension of a manifold, we put it in a superindex (i.e. MnM^{n} means that MM is a manifold of dimension nn) 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 f:MPf\colon M\to P is such that f1(P)=Mf^{-1}(\partial P)=\partial M and ff is transverse to the boundary P\partial P.

Definition 2.1.

Two smooth map germs

η:(n,0)(n+k,0)andϑ:(n,0)(n+k,0)\eta\colon(\mathbb{R}^{n},0)\to(\mathbb{R}^{n+k},0)\quad\text{and}\quad\vartheta\colon(\mathbb{R}^{n},0)\to(\mathbb{R}^{n+k},0)

are said to be left-right equivalent (in some sources also called 𝒜\mathscr{A}-equivalent) if there are diffeomorphism germs φ\varphi and ψ\psi of (n,0)(\mathbb{R}^{n},0) and (n+k,0)(\mathbb{R}^{n+k},0) respectively such that ϑ=ψηφ1\vartheta=\psi\circ\eta\circ\varphi^{-1}. The suspension of the germ η\eta is the germ

η×id1:(n+1,0)(n+k+1,0).\eta\times{\mathop{\rm id}}_{\mathbb{R}^{1}}\colon(\mathbb{R}^{n+1},0)\to(\mathbb{R}^{n+k+1},0).

By the singularity class (or simply singularity) of η\eta we mean the equivalence class of η\eta in the equivalence relation generated by left-right equivalence and suspension. The singularity class of η\eta is denoted by [η][\eta].

Observe that in each singularity class the codimension kk of the germs is fixed but the dimension nn is not.

Definition 2.2.

For manifolds Mn,Pn+kM^{n},P^{n+k} the product of the diffeomorphism groups of MM and PP has an action on the space C(M,P)C^{\infty}(M,P) of smooth maps f:MPf\colon M\to P defined by the formula (φ,ψ)ψfφ1(\varphi,\psi)\mapsto\psi\circ f\circ\varphi^{-1}. Two maps f,g:MPf,g\colon M\to P are said to be left-right equivalent if they are in the same orbit of this action, i.e. g=ψfφ1g=\psi\circ f\circ\varphi^{-1} for some diffeomorphisms φ\varphi of MM and ψ\psi of PP. A map f:MPf\colon M\to P is said to be stable if it is in the interior of an orbit, i.e. it has a neighbourhood UC(M,P)U\subset C^{\infty}(M,P) such that every element of UU is left-right equivalent to ff. 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 η\eta is stable, then every germ representing the singularity [η][\eta] 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 τ\tau of singularities of stable kk-codimensional germs.

Definition 2.4.

A smooth map f:MnPn+kf\colon M^{n}\to P^{n+k} is said to be a τ\tau-map if all of its germs belong to singularity classes in τ\tau. For a singularity class [η]τ[\eta]\in\tau a point pMp\in M is said to be an [η][\eta]-point if the germ of ff at pp is equivalent to η\eta; the set of [η][\eta]-points in MM is denoted by [η](f)[\eta](f). By a slight abuse of notation we will sometimes write η(f)\eta(f) instead of [η](f)[\eta](f) for better readability.

Definition 2.5.

For a singularity class [η][\eta] the minimal number mm for which there is a germ ϑ:(m,0)(m+k,0)\vartheta\colon(\mathbb{R}^{m},0)\to(\mathbb{R}^{m+k},0) in the singularity class [η][\eta] is called the codimension of the singularity [η][\eta] and is denoted by codim[η]\mathop{\rm codim}[\eta].

Remark 2.6.

If f:MPf\colon M\to P is a τ\tau-map, then MM is naturally stratified by the submanifolds η(f)\eta(f) for all [η]τ[\eta]\in\tau. Here η(f)\eta(f) is a submanifold of MM of codimension codim[η]\mathop{\rm codim}[\eta].

Example 2.7.
  1. (1)

    If τ\tau only contains the class Σ0\Sigma^{0} of regular germs (i.e. those with maximal-rank derivative at 0), then τ\tau-maps are just the immersions.

  2. (2)

    If τ\tau consists of Σ0\Sigma^{0} and all stable singularities of type Σ1\Sigma^{1}, that is, singularities with representatives whose derivative at 0 has corank 11, then τ\tau-maps are called Morin maps. Further restricting the set τ\tau of allowed singularities we can obtain e.g. the so-called fold maps (i.e. {Σ0,Σ1,0}\{\Sigma^{0},\Sigma^{1,0}\}-maps) and cusp maps (i.e. {Σ0,Σ1,0,Σ1,1,0}\{\Sigma^{0},\Sigma^{1,0},\Sigma^{1,1,0}\}-maps) where fold (Σ1,0\Sigma^{1,0}) and cusp (Σ1,1,0\Sigma^{1,1,0}) are the two simplest types of Morin singularities; see [mor].

In the following we shall work with τ\tau-maps to a fixed target manifold PP up to a cobordism relation defined as follows:

Definition 2.8.

We call two τ\tau-maps f0:M0nPn+kf_{0}\colon M_{0}^{n}\to P^{n+k} and f1:M1nPn+kf_{1}\colon M_{1}^{n}\to P^{n+k} (with closed source manifolds M0M_{0} and M1M_{1}) τ\tau-cobordant if there is

  • (i)

    a compact manifold Wn+1W^{n+1} with boundary such that W=M0M1\partial W=M_{0}\sqcup M_{1},

  • (ii)

    a τ\tau-map F:Wn+1P×[0,1]F\colon W^{n+1}\to P\times[0,1] such that for i=0,1i=0,1 we have F1(P×{i})=MiF^{-1}(P\times\{i\})=M_{i} and F|Mi=fiF|_{M_{i}}=f_{i}.

The τ\tau-cobordism class of f:MnPn+kf\colon M^{n}\to P^{n+k} is denoted by [f][f] and the set of all τ\tau-cobordism classes of τ\tau-maps to the manifold PP is denoted by Cobτ(P)\textstyle{\mathop{\rm Cob}}_{\tau}(P).

The set Cobτ(P)\textstyle{\mathop{\rm Cob}}_{\tau}(P) admits a natural commutative semigroup operation by the disjoint union: if f:MnPn+kf\colon M^{n}\to P^{n+k} and g:NnPn+kg\colon N^{n}\to P^{n+k} are τ\tau-maps, then so is

fg:MNPf\sqcup g\colon M\sqcup N\to P

and the τ\tau-cobordism class [f]+[g]:=[fg][f]+[g]:=[f\sqcup g] 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 Cobτ(P)\textstyle{\mathop{\rm Cob}}_{\tau}(P) is explicitly constructed in [hosszu]). This way Cobτ(P)\textstyle{\mathop{\rm Cob}}_{\tau}(P) becomes an Abelian group for any manifold PP.

Next we shall endow τ\tau-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 f:MPf\colon M\to P the stable isomorphism type of the virtual vector bundle νf:=fTPTM\nu_{f}:=f^{*}TP\ominus TM. If σ\sigma is such a decoration, then τ\tau-maps equipped with σ\sigma-structures will be denoted τσ\tau^{\sigma}-maps. In this paper we shall work with the following types of τσ\tau^{\sigma}-maps:

  1. (1)

    σ=G\sigma=G: Let GG be a stable group using the definition of Wall [stabgr, section 8.2] which we recall now:

    Definition 2.9.

    A stable group GG is defined as the direct limit of a sequence of group homomorphisms ιn:G(n)G(n+1)\iota_{n}\colon G(n)\to G(n+1) where

    • (i)

      there are homomorphisms αn:G(n)O(n)\alpha_{n}\colon G(n)\to\mathrm{O}(n) such that the diagram

      \textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιn1\scriptstyle{\iota_{n-1}}G(n)\textstyle{G(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αn\scriptstyle{\alpha_{n}}ιn\scriptstyle{\iota_{n}}G(n+1)\textstyle{G(n+1)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αn+1\scriptstyle{\alpha_{n+1}}ιn+1\scriptstyle{\iota_{n+1}}\textstyle{\ldots}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}O(n)\textstyle{\mathrm{O}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}O(n+1)\textstyle{\mathrm{O}(n+1)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots}

      is commutative where the inclusions in the lower row are natural,

    • (ii)

      there is a weakly increasing function c:c\colon\mathbb{N}\to\mathbb{N} tending to \infty such that the map ιn\iota_{n} is c(n)c(n)-connected,

    • (iii)

      there are homomorphisms βn,m:G(n)×G(m)G(n+m)\beta_{n,m}\colon G(n)\times G(m)\to G(n+m) such that the diagrams

      G(n)×G(m)\textstyle{G(n)\times G(m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιn×id\scriptstyle{\iota_{n}\times\mathop{\rm id}}βn,m\scriptstyle{\beta_{n,m}}G(n+m)\textstyle{G(n+m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιn+m\scriptstyle{\iota_{n+m}}G(n)×G(m)\textstyle{G(n)\times G(m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βn,m\scriptstyle{\beta_{n,m}}id×ιm\scriptstyle{\mathop{\rm id}\times\iota_{m}}G(n+1)×G(m)\textstyle{G(n+1)\times G(m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βn+1,m\scriptstyle{\beta_{n+1,m}}G(n+m+1)\textstyle{G(n+m+1)}G(n)×G(m+1)\textstyle{G(n)\times G(m+1)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βn,m+1\scriptstyle{\beta_{n,m+1}}
      G(n)×G(m)\textstyle{G(n)\times G(m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αn×αm\scriptstyle{\alpha_{n}\times\alpha_{m}}βn,m\scriptstyle{\beta_{n,m}}G(n+m)\textstyle{G(n+m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αn+m\scriptstyle{\alpha_{n+m}}O(n)×O(m)\textstyle{\mathrm{O}(n)\times\mathrm{O}(m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}O(n+m)\textstyle{\mathrm{O}(n+m)}

      and

      G(n)×G(m)×G(l)\textstyle{G(n)\times G(m)\times G(l)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βn,m×id\scriptstyle{\beta_{n,m}\times\mathop{\rm id}}id×βm,l\scriptstyle{\mathop{\rm id}\times\beta_{m,l}}G(n)×G(m+l)\textstyle{G(n)\times G(m+l)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βn,m+l\scriptstyle{\beta_{n,m+l}}G(n+m)×G(l)\textstyle{G(n+m)\times G(l)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βn+m,l\scriptstyle{\beta_{n+m,l}}G(n+m+l)\textstyle{G(n+m+l)}

      commute up to conjugation by an element in the component of the identity,

    • (iv)

      there is a commutative diagram

      G(n)×G(m)\textstyle{G(n)\times G(m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιn+mβn,m\scriptstyle{\iota_{n+m}\circ\beta_{n,m}}γn,m\scriptstyle{\gamma_{n,m}}G(m)×G(n)\textstyle{G(m)\times G(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιn+mβm,n\scriptstyle{\iota_{n+m}\circ\beta_{m,n}}O(n+m)\textstyle{\mathrm{O}(n+m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δn,m\scriptstyle{\delta_{n,m}}O(n+m)\textstyle{\mathrm{O}(n+m)}

      where γn,m\gamma_{n,m} denotes the interchange of factors and δn,m\delta_{n,m} is the conjugation by an element whose determinant has sign (1)nm(-1)^{nm}.

    We define τG\tau^{G}-maps to be τ\tau-maps f:MnPn+kf\colon M^{n}\to P^{n+k} for which the virtual normal bundle νf\nu_{f} has structure group GG in the following sense:

    Definition 2.10.

    For a stable group GG and a virtual vector bundle ν\nu we say that ν\nu has structure group GG if there is a fixed equivalence class of vector bundles stably isomorphic to ν\nu such that the structure group of each vector bundle ξ\xi is reduced to G(rkξ)G(\mathop{\rm rk}\xi) and their equivalence is defined as follows. Note that the reduction of the structure group of ξ\xi to G(rkξ)G(\mathop{\rm rk}\xi) also gives the reduction of the structure group of ξεr\xi\oplus\varepsilon^{r} to G(rkξ+r)G(\mathop{\rm rk}\xi+r) for any rr. Here and later on εr\varepsilon^{r} means the trivial rank-rr vector bundle over any base space. Now two such bundles ξ,ζ\xi,\zeta represetning ν\nu are said to be equivalent if for some rr the bundles ξεrrkξ\xi\oplus\varepsilon^{r-\mathop{\rm rk}\xi} and ζεrrkζ\zeta\oplus\varepsilon^{r-\mathop{\rm rk}\zeta} are isomorphic as G(r)G(r)-bundles.

    For example if G=SOG=\mathrm{SO}, then τSO\tau^{\mathrm{SO}}-maps are τ\tau-maps with oriented normal bundles; if G=OG=\mathrm{O}, then τO\tau^{\mathrm{O}}-maps are just τ\tau-maps without further conditions. If τ\tau consists of all possible singularities of kk-codimensional germs,then any manifold MnM^{n} has a τ\tau-map to n+k\mathbb{R}^{n+k} uniquely up to τ\tau-cobordism and its stable normal bundle is just the inverse of the tangent bundle TMTM in any KK-group of MM; the same is true for any abstract cobordism of manifolds, hence in this case we have

    CobτSO(n+k)=ΩnandCobτO(n+k)=𝔑n.\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(\mathbb{R}^{n+k})=\Omega_{n}\quad\text{and}\quad\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}}(\mathbb{R}^{n+k})=\mathfrak{N}_{n}.

    As another example, if τ={Σ0}\tau=\{\Sigma^{0}\} (i.e. τ\tau-maps are the immersions), then CobτG(n+k)\textstyle{\mathop{\rm Cob}}_{\tau}^{G}(\mathbb{R}^{n+k}) is the cobordism group of immersions of nn-manifolds into n+k\mathbb{R}^{n+k} with normal structure group GG which was first described by Wells [wells]. For various other singularity sets τ\tau with GG being mostly SO\mathrm{SO} or O\mathrm{O} the groups CobτG(P)\textstyle{\mathop{\rm Cob}}_{\tau}^{G}(P) were considered e.g. by Szűcs [analog], [hosszu]111We shall refer many times to [hosszu] which only considers τSO\tau^{\mathrm{SO}}-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 kk of the maps is non-positive, unlike in the present paper).

  2. (2)

    σ=int\sigma=\mathrm{int}: We define τint\tau^{\mathrm{int}}-maps to be τ\tau-maps with the first Stiefel–Whitney class being an integer class, that is, a τint\tau^{\mathrm{int}}-map is a pair (f,w)(f,w) where f:MnPn+kf\colon M^{n}\to P^{n+k} is a τ\tau-map and wH1(M;)w\in H^{1}(M;\mathbb{Z}) is a cohomology class such that its mod 22 reduction is w1(νf)w_{1}(\nu_{f}). Following e.g. Stong [stong] we call this type of normal structure a Wall structure and also call τint\tau^{\mathrm{int}}-maps Wall τ\tau-maps. Again, if τ\tau consists of all possible singularities of kk-codimensional germs, then any manifold MnM^{n} has a τ\tau-map to n+k\mathbb{R}^{n+k} uniquely up to τ\tau-cobordism and its first normal Stiefel–Whitney class is the same as w1(M)w_{1}(M), hence for this τ\tau we have

    Cobτint(n+k)=𝔚n.\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(\mathbb{R}^{n+k})=\mathfrak{W}_{n}.
  3. (3)

    σ=m\sigma=\oplus m for a natural number mm: τm\tau^{\oplus m}-maps are also called mm-framed τ\tau-maps and they are defined so that their germs are of the form η×idm\eta\times\mathop{\rm id}_{\mathbb{R}^{m}} (for [η]τ[\eta]\in\tau) and the change of coordinate neighbourhoods always induces the identity on m\mathbb{R}^{m} here. This extends the notion of mm-framed immersions, i.e. immersions equipped with mm pointwise independent normal vector fields; if τ={Σ0}\tau=\{\Sigma^{0}\}, then τm\tau^{\oplus m}-maps are just immersions with normal mm-framing. We can endow τm\tau^{\oplus m}-maps with other stable normal structures σ\sigma as well, and this defines τσm\tau^{\sigma\oplus m}-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 PP is of the form Q×mQ\times\mathbb{R}^{m} for a manifold QQ, then we have

    Cobτσm(Q×m)Cobτσ(Q)\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma\oplus m}(Q\times\mathbb{R}^{m})\cong\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma}(Q) (2.1)

    and the isomorphism is given by a natural correspondence between framed and non-framed maps; see [hosszu, proposition 13].

  4. (4)

    σ=Gξ\sigma=G\oplus\xi: Let GG be a stable group and ξ\xi a vector bundle over the classifying space BGBG. Then τGξ\tau^{G\oplus\xi}-maps are generalisations of τGm\tau^{G\oplus m}-maps and we postpone their precise definition to section 6. Intuitively a τGξ\tau^{G\oplus\xi}-map is a τG\tau^{G}-map such that a bundle induced from ξ\xi splits off from its “normal bundle” (the sense in which we mean this will be clarified in section 6); if ξ\xi is trivial of rank mm, this just gives an mm-framed τG\tau^{G}-map.

Now if σ\sigma is any of the above four stable normal structures, then the cobordism of two τσ\tau^{\sigma}-maps to a manifold PP can be defined by adding the σ\sigma-structure to the map FF in definition 2.8 and the cobordism group of τσ\tau^{\sigma}-maps to PP is denoted by Cobτσ(P)\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma}(P).

To form the statements of our main theorems it remains to define a vector bundle which will play the role of ξ\xi in a special type of τGξ\tau^{G\oplus\xi}-maps. We put G:=OG:=\mathrm{O} and denote by detγ\det\gamma the line bundle over BOB\mathrm{O} induced by the first Stiefel–Whitney class w1:BOPw_{1}\colon B\mathrm{O}\to\mathbb{R}P^{\infty} from the tautological line bundle. The notation of this line bundle is due to the fact that if γnO\gamma_{n}^{\mathrm{O}} denotes the tautological rank-nn vector bundle over BO(n)B\mathrm{O}(n), then the restriction of detγ\det\gamma over BO(n)B\mathrm{O}(n) is the determinant bundle detγnO\det\gamma_{n}^{\mathrm{O}}. In our second theorem we shall use the rank-22 bundle 2detγ=detγdetγ2\det\gamma=\det\gamma\oplus\det\gamma.

Now we are in a position to state our main results which are as follows:

Theorem I.  For any set τ\tau of stable singularities and any manifold QqQ^{q} there is a long exact sequence

\displaystyle\ldots ψm+1CobτSO(Q×m)φmCobτSO(Q×m)χmCobτint(Q×m)ψm\displaystyle\mathop{\xrightarrow{\leavevmode\nobreak\ \psi_{m+1}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \varphi_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}^{\mathrm{SO}}_{\tau}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \chi_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \psi_{m}\leavevmode\nobreak\ }}
ψmCobτSO(Q×m1)\displaystyle\mathop{\xrightarrow{\leavevmode\nobreak\ \psi_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m-1})\to\ldots

where φm\varphi_{m} is of the form id+ι\mathop{\rm id}+\iota where ι\iota is the involution represented by the reflection to a hyperplane, χm\chi_{m} is the forgetful homomorphism and ψm\psi_{m} assigns to a cobordism class [(f,w)][(f,w)] the class [f|PD(w)][f|_{\mathop{\rm PD}(w)}].

Theorem II.  For any set τ\tau of stable singularities and any manifold QqQ^{q} there is a long exact sequence

\displaystyle\ldots ψm+1Cobτint(Q×m)φmCobτO(Q×m)χmCobτO2detγ(Q×m)ψm\displaystyle\mathop{\xrightarrow{\leavevmode\nobreak\ \psi^{\prime}_{m+1}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \varphi^{\prime}_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}^{\mathrm{O}}_{\tau}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \chi^{\prime}_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \psi^{\prime}_{m}\leavevmode\nobreak\ }}
ψmCobτint(Q×m1)\displaystyle\mathop{\xrightarrow{\leavevmode\nobreak\ \psi^{\prime}_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m-1})\to\ldots

where φm\varphi^{\prime}_{m} is the forgetful homomorphism and χm\chi^{\prime}_{m} assigns to a cobordism class [f][f] the class [f|PD(w1(νf)2)][f|_{\mathop{\rm PD}(w_{1}(\nu_{f})^{2})}].

Remark 2.11.

We shall give a geometric description of ψm\psi^{\prime}_{m} later in remark 8.3. Here we just note that its classical analogue is always zero while in our general case this does not always hold; proposition 9.3 will show that even the composition ψm1ψm\psi_{m-1}\circ\psi^{\prime}_{m} is not necessarily zero.

Remark 2.12.

Although we made assumptions on the set τ\tau of allowed singularities which generally exclude the set of all singularities, i.e. bordism groups in general cannot be considered as τ\tau-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 QMSO(Q)Q\mapsto M\mathrm{SO}^{*}(Q); see [atiyah]. Similarly cobordism groups of singular maps also yield cohomology theories: for the singularity sets τ\tau and stable normal structures σ\sigma used in the present paper we can define extraordinary cohomology functors hτσh^{*}_{\tau^{\sigma}} (see [hosszu, section 19]) and if QQ is a manifold, then we have

hτσm(Q)=Cobτσ(Q×m)andhτσm(Q)=Cobτσm(Q)h^{-m}_{\tau^{\sigma}}(Q)=\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma}(Q\times\mathbb{R}^{m})\quad\text{and}\quad h^{m}_{\tau^{\sigma}}(Q)=\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma\oplus m}(Q)

for any natural number mm (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 Cobτσm(Q)\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma\oplus-m}(Q) in the place of Cobτσ(Q×m)\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma}(Q\times\mathbb{R}^{m}) for negative mm’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 ΩnΩn\Omega_{n}\to\Omega_{n} is the multiplication by 22, hence for theorem 2 to be completely analogous to it we should have φm=2id\varphi_{m}=2\mathop{\rm id}, i.e. ι=id\iota=\mathop{\rm id}, which is indeed so in some special cases but not generally as we shall see. Moreover, ψm\psi_{m} is [(f,w)][f|PD(w)][(f,w)]\mapsto[f|_{\mathop{\rm PD}(w)}] and not [(f,w)][f|PD(w1(νf))][(f,w)]\mapsto[f|_{\mathop{\rm PD}(w_{1}(\nu_{f}))}] although its classical version 𝔚nΩn1\mathfrak{W}_{n}\to\Omega_{n-1} in (I) is just [M][PD(w1(M))][M]\mapsto[\mathop{\rm PD}(w_{1}(M))]. This is because 𝔚n\mathfrak{W}_{n} embeds into 𝔑n\mathfrak{N}_{n}, hence taking the Poincaré dual of the integer first Stiefel–Whitney class of a Wall manifold MM up to cobordism is independent of which integer representative of w1(M)w_{1}(M) we take, however, remark 2.11 implies that φm\varphi^{\prime}_{m} 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 O2detγ\mathrm{O}\oplus 2\det\gamma-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 τint\tau^{\mathrm{int}}-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 Cobτσ(P)\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma}(P) (where τ\tau is a set of stable singularities, σ\sigma is a stable normal structure and PP is an arbitrary manifold) will be denoted by XτσX_{\tau}^{\sigma}, that is, XτσX_{\tau}^{\sigma} is the (homotopically unique) space which has the property

Cobτσ(P)[𝑃,Xτσ]\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma}(P)\cong[\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},X_{\tau}^{\sigma}]

where 𝑃\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}} denotes the one-point compactification of PP and [,][\cdot,\cdot] means the (based) homotopy classes of maps from the first space to the second.

If σ=G\sigma=G is a stable group, then many explicit constructions of the space XτGX_{\tau}^{G} 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 XτintX_{\tau}^{\mathrm{int}} of τint\tau^{\mathrm{int}}-cobordisms.

A stable group GG is defined as the direct limit of a sequence of groups G(n)G(n) where G(n)G(n) has a fixed linear action on n\mathbb{R}^{n} (see definition 2.9). We denote by γnG\gamma_{n}^{G} the universal rank-nn vector bundle with structure group GG, i.e.

γnG:=EG(n)×G(n)n\gamma_{n}^{G}:=EG(n)\mathop{\mathop{\times}\limits_{\vbox to-1.50694pt{\kern-5.425pt\hbox{$\scriptstyle G(n)$}\vss}}}\mathbb{R}^{n}

where EG(n)G(n)BG(n)EG(n)\mathop{\xrightarrow{\leavevmode\nobreak\ G(n)\leavevmode\nobreak\ }}BG(n) is the universal principal G(n)G(n)-bundle.

Definition 3.1.

As before, let τ\tau be a set of stable singularities and GG a stable group. Consider the jet bundle J0(εn,γn+kG)J_{0}^{\infty}(\varepsilon^{n},\gamma_{n+k}^{G}) over BG(n+k)BG(n+k) which has as fibre the infinite jet space J0(n,n+k)J_{0}^{\infty}(\mathbb{R}^{n},\mathbb{R}^{n+k}) (i.e. the space of all polynomial maps nn+k\mathbb{R}^{n}\to\mathbb{R}^{n+k} with 0 constant term). Denote by Vτ(n)J0(n,n+k)V_{\tau}(n)\subset J_{0}^{\infty}(\mathbb{R}^{n},\mathbb{R}^{n+k}) the subspace of those maps whose singularity at 0 is in τ\tau and let KτG(n)K_{\tau}^{G}(n) be the union of the Vτ(n)V_{\tau}(n)’s in each fibre of J0(εn,γn+kG)J_{0}^{\infty}(\varepsilon^{n},\gamma_{n+k}^{G}), that is,

KτG(n):=EG(n+k)×G(n+k)Vτ(n)K_{\tau}^{G}(n):=EG(n+k)\mathop{\mathop{\times}\limits_{\vbox to-1.50694pt{\kern-5.425pt\hbox{$\scriptstyle G(n+k)$}\vss}}}V_{\tau}(n)

where the action of G(n+k)G(n+k) on Vτ(n)V_{\tau}(n) is one-sided, we do not act on the source space n\mathbb{R}^{n} of the polynomial maps. Now the natural map G(n+k)G(n+k+1)G(n+k)\to G(n+k+1) gives us a map KτG(n)KτG(n+1)K_{\tau}^{G}(n)\to K_{\tau}^{G}(n+1). The Kazarian space for τG\tau^{G}-maps is the direct limit

KτG:=limnKτG(n).K_{\tau}^{G}:=\lim_{n\to\infty}K_{\tau}^{G}(n).

Note that we have a fibration KτG(n)Vτ(n)BG(n+k)K_{\tau}^{G}(n)\mathop{\xrightarrow{\leavevmode\nobreak\ V_{\tau}(n)\leavevmode\nobreak\ }}BG(n+k).

Definition 3.2.

The pullback of the vector bundle γn+kGBG(n+k)\gamma_{n+k}^{G}\to BG(n+k) by the above fibration will be denoted by ντG(n)εnKτG(n)\nu_{\tau}^{G}(n)\oplus\varepsilon^{n}\to K_{\tau}^{G}(n) where ντG(n)\nu_{\tau}^{G}(n) is a rank-kk virtual vector bundle. Since the natural map BG(n+k)BG(n+k+1)BG(n+k)\to BG(n+k+1) induces γn+kGε1\gamma_{n+k}^{G}\oplus\varepsilon^{1} from γn+k+1G\gamma_{n+k+1}^{G}, we get that the map KτG(n)KτG(n+1)K_{\tau}^{G}(n)\to K_{\tau}^{G}(n+1) induces ντG(n)\nu_{\tau}^{G}(n) from ντG(n+1)\nu_{\tau}^{G}(n+1). The universal virtual normal bundle of τG\tau^{G}-maps is

ντG:=limnντG(n)KτG.\nu_{\tau}^{G}:=\lim_{n\to\infty}\nu_{\tau}^{G}(n)\to K_{\tau}^{G}.

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 τG\tau^{G}-cobordisms this is the strengthened Kazarian conjecture [hosszu, corollary 74] and for τint\tau^{\mathrm{int}}- and τGξ\tau^{G\oplus\xi}-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 Γ:=ΩS\Gamma:=\Omega^{\infty}S^{\infty} 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 ν\nu be a virtual vector bundle of rank kk over a space KK such that KK is the direct limit of a sequence of inclusions K(n)K(n+1)K(n)\subset K(n+1) where for each nn the restriction ν(n):=ν|K(n)\nu(n):=\nu|_{K(n)} can be represented by ξ(n)εn\xi(n)\ominus\varepsilon^{n} for a (non-virtual) vector bundle ξ(n)\xi(n), moreover, the inclusion K(n)K(n+1)K(n)\hookrightarrow K(n+1) is a c(n)c(n)-homotopy equivalence for a weakly increasing sequence c:c\colon\mathbb{N}\to\mathbb{N} tending to \infty (this holds for ντG\nu_{\tau}^{G} but also more generally for any virtual bundle over a CW complex). Then we can define SnTν(n)S^{n}T\nu(n) as the Thom space Tξ(n)T\xi(n), hence ΓTν(n)=ΩnΓSnTν(n)\Gamma T\nu(n)=\Omega^{n}\Gamma S^{n}T\nu(n) also exists. We define the space

ΓTν:=limnΓTν(n)=limnlimmΩn+mSn+mTν(n).\Gamma T\nu:=\lim_{n\to\infty}\Gamma T\nu(n)=\lim_{n\to\infty}\lim_{m\to\infty}\Omega^{n+m}S^{n+m}T\nu(n).

Note that here the inclusion Sn+mTν(n)Sn+mTν(n+1)S^{n+m}T\nu(n)\hookrightarrow S^{n+m}T\nu(n+1) induces isomorphism in cohomologies up to the index c(n)+k+n+mc(n)+k+n+m for each m>0m>0 (by the Thom isomorphism), hence the (c(n)+k)(c(n)+k)-homotopy type of Ωn+mSn+mν(n)\Omega^{n+m}S^{n+m}\nu(n) coincides with that of Ωn+mSn+mTν(n+1)\Omega^{n+m}S^{n+m}T\nu(n+1) (by the Whitehead theorem). This means that the (c(n)+k)(c(n)+k)-homotopy type of Ωn+mSn+mTν\Omega^{n+m}S^{n+m}T\nu can be defined as that of Ωn+mSn+mTν(n)\Omega^{n+m}S^{n+m}T\nu(n). Moreover, we also get that the direct limits ΓTν(n)\Gamma T\nu(n) and ΓTν(n+1)\Gamma T\nu(n+1) are (c(n)+k)(c(n)+k)-homotopy equivalent and since c(n)c(n) tends to \infty this implies that for any rr the rr-homotopy type of ΓTν\Gamma T\nu is that of Ωn+mSn+mTν(n)\Omega^{n+m}S^{n+m}T\nu(n) for some numbers nn and mm.

Remark 3.4.

The space ΓTν\Gamma T\nu is an infinite loop space, hence it naturally defines a spectrum E:=ΩSTν=ΓSTνE_{*}:=\Omega^{\infty-*}S^{\infty}T\nu=\Gamma S^{*}T\nu. What is more, the “virtual space” TνT\nu defines an equivalence class of spectra and this EE_{*} is a representative of it; see [hosszu, remark 62].

The strengthened Kazarian conjecture [hosszu, corollary 74] states:

Theorem 3.5.

If τ\tau is a set of stable singularities and GG is a stable group, then the classifying space XτGX_{\tau}^{G} of cobordisms of τG\tau^{G}-maps is homotopy equivalent to ΓTντG\Gamma T\nu_{\tau}^{G}.

Remark 3.6.

The cohomology theory mentioned in remark 2.13 is defined for σ=G\sigma=G by the spectrum EτG:=ΓSTντGE_{*}^{\tau^{G}}:=\Gamma S^{*}T\nu_{\tau}^{G}.

Remark 3.7.

Later it will be important for us to have a deeper understanding of the connection between the cobordism group CobτG(P)\textstyle{\mathop{\rm Cob}}_{\tau}^{G}(P) (for a manifold Pn+kP^{n+k}) and the classifying homotopy classes [𝑃,ΓTντG][\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Gamma T\nu_{\tau}^{G}], so we now give a short description of it based on [hosszu].

If f:MnPn+kf\colon M^{n}\to P^{n+k} is a τG\tau^{G}-map and i:Mnri\colon M^{n}\hookrightarrow\mathbb{R}^{r} is an embedding, then f×i:MP×rf\times i\colon M\hookrightarrow P\times\mathbb{R}^{r} is a so-called τG\tau^{G}-embedding. A τG\tau^{G}-embedding of MnM^{n} into a manifold Qn+k+rQ^{n+k+r} is defined as a triple (e,𝒱,)(e,\mathscr{V},\mathscr{F}) where e:MQe\colon M\hookrightarrow Q is an embedding with normal GG-structure, 𝒱\mathscr{V} is a sequence (v1,,vr)(v_{1},\ldots,v_{r}) where the viv_{i}’s are pointwise independent vector fields along e(M)e(M) (i.e. sections of the bundle TQ|e(M)TQ|_{e(M)}), \mathscr{F} is a foliation of dimension rr on a neighbourhood of e(M)e(M) which is tangent to 𝒱\mathscr{V} along e(M)e(M) and any point pMp\in M has a neighbourhood on which the composition of ee with the projection along the leaves of \mathscr{F} to a small (n+k)(n+k)-dimensional transverse slice has at pp a singularity which belongs to τ\tau.

Cobordisms of τG\tau^{G}-embeddings of nn-manifolds to QQ can be defined in the usual way and their cobordism group is denoted by EmbτG(n,Q)\textstyle{\mathop{\rm Emb}}_{\tau}^{G}(n,Q). Now if the number rr is sufficiently large, then assigning to the τG\tau^{G}-map f:MnPn+kf\colon M^{n}\to P^{n+k} the τG\tau^{G}-embedding f×i:MP×rf\times i\colon M\hookrightarrow P\times\mathbb{R}^{r} (with vector fields arising from a basis of r\mathbb{R}^{r} and foliation composed of the leaves {p}×r\{p\}\times\mathbb{R}^{r}) yields an isomorphism

CobτG(P)EmbτG(n,P×r).\textstyle{\mathop{\rm Cob}}_{\tau}^{G}(P)\cong\textstyle{\mathop{\rm Emb}}_{\tau}^{G}(n,P\times\mathbb{R}^{r}).

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 P×rP\times\mathbb{R}^{r} which leave the stable normal bundles of the maps involved unchanged.

Now forgetting the singularity structure (given by 𝒱\mathscr{V} and \mathscr{F}) of f×if\times i yields a well-defined cobordism class of an embedding of MM into P×rP\times\mathbb{R}^{r} with normal GG-structure. Since these are classified by the Thom space Tγk+rGT\gamma_{k+r}^{G} we have the homotopy class of a map Sr𝑃Tγk+rGS^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\to T\gamma_{k+r}^{G} corresponding to it such that f×i(M)f\times i(M) is the preimage of BG(k+r)BG(k+r) and its normal bundle is induced from γr+kG\gamma_{r+k}^{G} (note that Sr𝑃S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}} is the one-point compactification of P×rP\times\mathbb{R}^{r}). Then using the singularity structure we can lift this classifying map through the map T(ντG(r)εr)Tγk+rGT(\nu_{\tau}^{G}(r)\oplus\varepsilon^{r})\to T\gamma_{k+r}^{G} induced by the fibration KτG(r)BG(k+r)K_{\tau}^{G}(r)\to BG(k+r) (see definition 3.2) to get a map

κf:Sr𝑃T(ντG(r)εr)SrTντG(r)\kappa_{f}\colon S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\to T(\nu_{\tau}^{G}(r)\oplus\varepsilon^{r})\cong S^{r}T\nu_{\tau}^{G}(r)

where f×i(M)f\times i(M) is now the preimage of KτG(r)K_{\tau}^{G}(r) and its normal bundle is induced from ντG(r)εr\nu_{\tau}^{G}(r)\oplus\varepsilon^{r}.

The adjoint correspondence identifies [Sr𝑃,SrTντG(r)][S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T\nu_{\tau}^{G}(r)] with [𝑃,ΩrSrTντG(r)][\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Omega^{r}S^{r}T\nu_{\tau}^{G}(r)] and if rr is sufficiently large, the latter is identified with [𝑃,ΩSTντG(r)][\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Omega^{\infty}S^{\infty}T\nu_{\tau}^{G}(r)] since by increasing rr we only attach large dimensional cells to the target space. Now the homotopy class [κf][\kappa_{f}] can be thought of as an element in [𝑃,ΓTντG][\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Gamma T\nu_{\tau}^{G}] and assigning [κf][\kappa_{f}] to [f][f] gives an isomorphism

EmbτG(n,P×r)[𝑃,ΓTντG].\textstyle{\mathop{\rm Emb}}_{\tau}^{G}(n,P\times\mathbb{R}^{r})\cong[\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Gamma T\nu_{\tau}^{G}].

The composition of the two isomorphisms described above show a bijective correspondence between the cobordism classes of the ff’s and the homotopy classes of the κf\kappa_{f}’s which forms the isomorphism CobτG(P)[𝑃,ΓTντG]\textstyle{\mathop{\rm Cob}}_{\tau}^{G}(P)\cong[\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Gamma T\nu_{\tau}^{G}] of theorem 3.5.

Next we shall define the Kazarian space of Wall τ\tau-maps and describe the classifying space XτintX_{\tau}^{\mathrm{int}} similarly to the theorem above. Consider the first Stiefel–Whitney class w1:BOPw_{1}\colon B\mathrm{O}\to\mathbb{R}P^{\infty} (which is a fibration with fibre BSOB\mathrm{SO}) and let i:S1Pi\colon S^{1}\hookrightarrow\mathbb{R}P^{\infty} be the inclusion of the 11-cell.

Definition 3.8.

Let BintB\mathrm{int} be the homotopy pullback of the diagram BOw1P𝑖S1B\mathrm{O}\mathop{\xrightarrow{\leavevmode\nobreak\ w_{1}\leavevmode\nobreak\ }}\mathbb{R}P^{\infty}\xleftarrow{i}S^{1}, i.e. the space we obtain by pulling back the fibration w1:BOBSOPw_{1}\colon B\mathrm{O}\mathop{\xrightarrow{\leavevmode\nobreak\ B\mathrm{SO}\leavevmode\nobreak\ }}\mathbb{R}P^{\infty} by ii, and let w:BintS1w\colon B\mathrm{int}\to S^{1} be the map indicated on the pullback diagram

Bint\textstyle{B\mathrm{int}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w\scriptstyle{w}S1\textstyle{S^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{i}BO\textstyle{B\mathrm{O}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w1\scriptstyle{w_{1}}P\textstyle{\mathbb{R}P^{\infty}}

In other words we define BintB\mathrm{int} as the subspace S1×2BSOS^{1}\mathop{\mathop{\times}\limits_{\vbox to-1.50694pt{\kern-5.425pt\hbox{$\scriptstyle\mathbb{Z}_{2}$}\vss}}}B\mathrm{SO} in BO=S×2BSOB\mathrm{O}=S^{\infty}\mathop{\mathop{\times}\limits_{\vbox to-1.50694pt{\kern-5.425pt\hbox{$\scriptstyle\mathbb{Z}_{2}$}\vss}}}B\mathrm{SO}.

This BintB\mathrm{int} is the direct limit of the spaces Bint(n)B\mathrm{int}(n) that we get in a similar way from BO(n)B\mathrm{O}(n) instead of BOB\mathrm{O}. We can pull back the tautological bundle γnOBO(n)\gamma_{n}^{\mathrm{O}}\to B\mathrm{O}(n) to a vector bundle γnintBint(n)\gamma_{n}^{\mathrm{int}}\to B\mathrm{int}(n) and this makes Bint(n)B\mathrm{int}(n) the classifying space of rank-nn vector bundles equipped with Wall structures, that is, those vector bundles ξ\xi for which w1(ξ)w_{1}(\xi) is the mod 22 reduction of a specific integer cohomology class (this is true since S1K(,1)S^{1}\cong K(\mathbb{Z},1), PK(2,1)\mathbb{R}P^{\infty}\cong K(\mathbb{Z}_{2},1) and ii corresponds to the mod 22 reduction). Now we can again consider the jet bundle J0(εn,γn+kint)J_{0}^{\infty}(\varepsilon^{n},\gamma_{n+k}^{\mathrm{int}}) over Bint(n+k)B\mathrm{int}(n+k) and take in each fibre the space Vτ(n)V_{\tau}(n) to obtain a fibration Kτint(n)Vτ(n)Bint(n+k)K_{\tau}^{\mathrm{int}}(n)\mathop{\xrightarrow{\leavevmode\nobreak\ V_{\tau}(n)\leavevmode\nobreak\ }}B\mathrm{int}(n+k) as in definition 3.1.

Definition 3.9.

If τ\tau is a set of stable singularities, the Kazarian space for τint\tau^{\mathrm{int}}-maps is Kτint:=limnKτint(n)K_{\tau}^{\mathrm{int}}:=\mathop{\lim}\limits_{\vbox to-1.50694pt{\kern-4.82224pt\hbox{$\scriptstyle n\to\infty$}\vss}}K_{\tau}^{\mathrm{int}}(n).

We have now a diagram

Kτint(n)\textstyle{K_{\tau}^{\mathrm{int}}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Vτ(n)\scriptstyle{V_{\tau}(n)}KτO(n)\textstyle{K_{\tau}^{\mathrm{O}}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Vτ(n)\scriptstyle{V_{\tau}(n)}J0(εn,γn+kint)\textstyle{J^{\infty}_{0}(\varepsilon^{n},\gamma^{\mathrm{int}}_{n+k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}J0(n,n+k)\scriptstyle{J^{\infty}_{0}(\mathbb{R}^{n},\mathbb{R}^{n+k})}J0(εn,γn+kO)\textstyle{J^{\infty}_{0}(\varepsilon^{n},\gamma^{\mathrm{O}}_{n+k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}J0(n,n+k)\scriptstyle{J^{\infty}_{0}(\mathbb{R}^{n},\mathbb{R}^{n+k})}Bint(n+k)\textstyle{B\mathrm{int}(n+k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BO(n+k)\textstyle{B\mathrm{O}(n+k)}

where all squares are pullback squares. Using this we can define the rank-kk virtual vector bundle ντint(n)Kτint(n)\nu_{\tau}^{\mathrm{int}}(n)\to K_{\tau}^{\mathrm{int}}(n) as the pullback of ντO(n)KτO(n)\nu_{\tau}^{\mathrm{O}}(n)\to K_{\tau}^{\mathrm{O}}(n).

Definition 3.10.

The universal virtual normal bundle of τint\tau^{\mathrm{int}}-maps is ντint:=limnντint(n)\nu_{\tau}^{\mathrm{int}}:=\mathop{\lim}\limits_{\vbox to-1.50694pt{\kern-4.82224pt\hbox{$\scriptstyle n\to\infty$}\vss}}\nu_{\tau}^{\mathrm{int}}(n) over KτintK_{\tau}^{\mathrm{int}}.

Theorem 3.11.

If τ\tau is a set of stable singularities, then we have XτintΓTντintX_{\tau}^{\mathrm{int}}\cong\Gamma T\nu_{\tau}^{\mathrm{int}}.

Proof.  We need an isomorphism Cobτint(P)[𝑃,ΓTντint]\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(P)\cong[\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Gamma T\nu_{\tau}^{\mathrm{int}}] for any target PP. Recall from remark 3.7 that the analogous isomorphism for τO\tau^{\mathrm{O}}-cobordisms is obtained by assigning τ\tau-embeddings to the τ\tau-maps, then lifting their classifying maps through the map SrTντO(r)Tγr+kOS^{r}T\nu_{\tau}^{\mathrm{O}}(r)\to T\gamma_{r+k}^{\mathrm{O}} (where rr is a large number). As we noted in remark 3.7 the isomorphism of the cobordism groups of τ\tau-maps and τ\tau-embeddings (see [hosszu, theorem 2]) does not depend on the choice of normal structures, hence it also yields the isomorphism

Cobτint(P)Embτint(n,P×r)\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(P)\cong\textstyle{\mathop{\rm Emb}}_{\tau}^{\mathrm{int}}(n,P\times\mathbb{R}^{r})

for rr sufficiently large, where Embτint(n,P×r)\textstyle{\mathop{\rm Emb}}_{\tau}^{\mathrm{int}}(n,P\times\mathbb{R}^{r}) is defined analogously.

Now if we have an embedding with normal Wall structure, then its classifying map can be lifted to Tγr+kintT\gamma_{r+k}^{\mathrm{int}} so if it is also a τ\tau-embedding, then using the pullback square

SrTντint(r)\textstyle{S^{r}T\nu_{\tau}^{\mathrm{int}}(r)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SrTντO(r)\textstyle{S^{r}T\nu_{\tau}^{\mathrm{O}}(r)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tγr+kint\textstyle{T\gamma_{r+k}^{\mathrm{int}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tγr+kO\textstyle{T\gamma_{r+k}^{\mathrm{O}}}

this inducing map lifts to SrTντint(r)S^{r}T\nu_{\tau}^{\mathrm{int}}(r). So τint\tau^{\mathrm{int}}-embeddings can be induced from SrTντint(r)S^{r}T\nu_{\tau}^{\mathrm{int}}(r) and since all of the above extnds to maps between manifolds with boundary (in particular to cobordisms), we get a homomorphism from Cobτint(P)\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(P) to [𝑃,ΓTντint][\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Gamma T\nu_{\tau}^{\mathrm{int}}] as in remark 3.7.

To see that this is an isomorphism, note that if we have a map Sr𝑃SrTντint(r)S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\to S^{r}T\nu_{\tau}^{\mathrm{int}}(r), then its composition with the map to SrTντO(r)S^{r}T\nu_{\tau}^{\mathrm{O}}(r) pulls back a τ\tau-embedding into P×rP\times\mathbb{R}^{r} (by theorem 3.5) which also has a normal Wall structure. Similarly homotopies between maps to SrTντint(r)S^{r}T\nu_{\tau}^{\mathrm{int}}(r) pull back cobordisms of τ\tau-embeddings with normal Wall structures. Thus we obtained the inverse of the above homomorphism Cobτint(P)[𝑃,ΓTντint]\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(P)\to[\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Gamma T\nu_{\tau}^{\mathrm{int}}], meaning that it is iso and this concludes our proof.  \square

4 A long exact sequence

We shall now connect the classifying spaces of τint\tau^{\mathrm{int}}-cobordisms and τSO\tau^{\mathrm{SO}}-cobordisms. Consider the embedding BSOBOB\mathrm{SO}\hookrightarrow B\mathrm{O} as a fibre over a point in S1PS^{1}\subset\mathbb{R}P^{\infty} which then factors through v:BSOBintv\colon B\mathrm{SO}\hookrightarrow B\mathrm{int} (again as the embedding of a fibre):

Bint\textstyle{B\mathrm{int}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S1×2BSO\textstyle{S^{1}\mathop{\mathop{\times}\limits_{\vbox to-1.50694pt{\kern-5.425pt\hbox{$\scriptstyle\mathbb{Z}_{2}$}\vss}}}B\mathrm{SO}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w\scriptstyle{w}S1\textstyle{S^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{i}BSO\textstyle{B\mathrm{SO}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v\scriptstyle{v}BO\textstyle{B\mathrm{O}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S×2BSO\textstyle{S^{\infty}\mathop{\mathop{\times}\limits_{\vbox to-1.50694pt{\kern-5.425pt\hbox{$\scriptstyle\mathbb{Z}_{2}$}\vss}}}B\mathrm{SO}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w1\scriptstyle{w_{1}}P\textstyle{\mathbb{R}P^{\infty}}

If we now take the approximations BSO(n)B\mathrm{SO}(n), Bint(n)B\mathrm{int}(n) and BO(n)B\mathrm{O}(n) of the spaces above and the approximation v(n):BSO(n)Bint(n)v(n)\colon B\mathrm{SO}(n)\to B\mathrm{int}(n) of vv, then both squares in the diagram below will be pullback squares:

γnSO\textstyle{\gamma^{\mathrm{SO}}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γnint\textstyle{\gamma^{\mathrm{int}}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γnO\textstyle{\gamma^{\mathrm{O}}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BSO(n)\textstyle{B\mathrm{SO}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v(n)\scriptstyle{v(n)}Bint(n)\textstyle{B\mathrm{int}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BO(n)\textstyle{B\mathrm{O}(n)}

But then the same is true for the jet bundles over the spaces BSO(n+k)B\mathrm{SO}(n+k), Bint(n+k)B\mathrm{int}(n+k) and BO(n+k)B\mathrm{O}(n+k), hence also for the Kazarian spaces and so the suspensions of the approximations of the universal virtual normal bundles also fit into a diagram

ντSO(n)εn\textstyle{\nu_{\tau}^{\mathrm{SO}}(n)\oplus\varepsilon^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ντint(n)εn\textstyle{\nu_{\tau}^{\mathrm{int}}(n)\oplus\varepsilon^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ντO(n)εn\textstyle{\nu_{\tau}^{\mathrm{O}}(n)\oplus\varepsilon^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γn+kSO\textstyle{\gamma^{\mathrm{SO}}_{n+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γn+kint\textstyle{\gamma^{\mathrm{int}}_{n+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γn+kO\textstyle{\gamma^{\mathrm{O}}_{n+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KτSO(n)\textstyle{K_{\tau}^{\mathrm{SO}}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kτint(n)\textstyle{K_{\tau}^{\mathrm{int}}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KτO(n)\textstyle{K_{\tau}^{\mathrm{O}}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BSO(n+k)\textstyle{B\mathrm{SO}(n+k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v(n+k)\scriptstyle{v(n+k)}Bint(n+k)\textstyle{B\mathrm{int}(n+k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BO(n+k)\textstyle{B\mathrm{O}(n+k)}

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 τ\tau of stable singularities and any manifold QqQ^{q} there is a long exact sequence

\displaystyle\ldots ψm+1CobτSO(Q×m)φmCobτSO(Q×m)χmCobτint(Q×m)ψm\displaystyle\mathop{\xrightarrow{\leavevmode\nobreak\ \psi_{m+1}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \varphi_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}^{\mathrm{SO}}_{\tau}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \chi_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \psi_{m}\leavevmode\nobreak\ }}
ψmCobτSO(Q×m1)\displaystyle\mathop{\xrightarrow{\leavevmode\nobreak\ \psi_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m-1})\to\ldots

Proof.  Let AS1A\subset S^{1} be any contractible subspace and let B:=w1(A)B:=w^{-1}(A) be its preimage in BintB\mathrm{int}. If v:BSOBintv\colon B\mathrm{SO}\hookrightarrow B\mathrm{int} was the embedding of the fibre over a point in AA, then v(BSO)Bv(B\mathrm{SO})\subset B and the map vv is a homotopy equivalence BSOBB\mathrm{SO}\cong B. Moreover, if KK is the restriction of the Kazarian space KτintK_{\tau}^{\mathrm{int}} over BBintB\subset B\mathrm{int}, then we have a pullback square

KτSO\textstyle{K_{\tau}^{\mathrm{SO}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kτint|B=K\textstyle{K_{\tau}^{\mathrm{int}}|_{B}=K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BSO\textstyle{B\mathrm{SO}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}B\textstyle{B}

with the vertical arrows being fibrations, hence we have KKτSOK\cong K_{\tau}^{\mathrm{SO}}. Similarly if ν\nu is the restriction of ντint\nu_{\tau}^{\mathrm{int}} over KKτintK\subset K_{\tau}^{\mathrm{int}}, then we have a pullback square

ντSO\textstyle{\nu_{\tau}^{\mathrm{SO}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ντint|K=ν\textstyle{\nu_{\tau}^{\mathrm{int}}|_{K}=\nu\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KτSO\textstyle{K_{\tau}^{\mathrm{SO}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}K\textstyle{K}

which implies that ν\nu is stably isomorphic to ντSO\nu_{\tau}^{\mathrm{SO}}.

Now let A1A_{1} be any point pS1p\in S^{1} and put A2:=S1{p}A_{2}:=S^{1}\setminus\{p\} (which are contractible subspaces). Let the preimages of A1A_{1} and A2A_{2} under ww respectively be B1B_{1} and B2B_{2} (thus we have B1BSOB2B_{1}\cong B\mathrm{SO}\cong B_{2}), their preimages in KτintK_{\tau}^{\mathrm{int}} be K1K_{1} and K2K_{2} (thus we have K1KτSOK2K_{1}\cong K_{\tau}^{\mathrm{SO}}\cong K_{2}) and put ν1:=ντint|K1\nu_{1}:=\nu_{\tau}^{\mathrm{int}}|_{K_{1}} and ν2:=ντint|K2\nu_{2}:=\nu_{\tau}^{\mathrm{int}}|_{K_{2}} (which are now stably isomorphic to ντSO\nu_{\tau}^{\mathrm{SO}}). Then for all nn we have a cofibraion

SnTν2(n)SnTντint(n)SnTντint(n)/SnTν2(n)S^{n}T\nu_{2}(n)\hookrightarrow S^{n}T\nu_{\tau}^{\mathrm{int}}(n)\to S^{n}T\nu_{\tau}^{\mathrm{int}}(n)/S^{n}T\nu_{2}(n)

(where ν1(n)\nu_{1}(n) and ν2(n)\nu_{2}(n) are the appropriate restrictions of ντint(n)\nu_{\tau}^{\mathrm{int}}(n)). Since the normal bundle of K1KτintK_{1}\subset K_{\tau}^{\mathrm{int}} is induced from the normal bundle of pS1p\in S^{1} it is trivial and so we have SnTντint(n)/SnTν2(n)=Sn+1Tν1(n)S^{n}T\nu_{\tau}^{\mathrm{int}}(n)/S^{n}T\nu_{2}(n)=S^{n+1}T\nu_{1}(n), hence the cofibration above has the form

SnTντSO(n)SnTντint(n)Sn+1TντSO(n).S^{n}T\nu_{\tau}^{\mathrm{SO}}(n)\hookrightarrow S^{n}T\nu_{\tau}^{\mathrm{int}}(n)\to S^{n+1}T\nu_{\tau}^{\mathrm{SO}}(n).

Now applying the functor Ωn+mΓ\Omega^{n+m}\Gamma to the Puppe sequence of this cofibration we get a sequence of maps

ΩmΓTντSO(n)ΩmΓTντint(n)ΩmΓSTντSO(n)Ωm1ΓTντSO(n)\ldots\to\Omega^{m}\Gamma T\nu_{\tau}^{\mathrm{SO}}(n)\to\Omega^{m}\Gamma T\nu_{\tau}^{\mathrm{int}}(n)\to\Omega^{m}\Gamma ST\nu_{\tau}^{\mathrm{SO}}(n)\to\Omega^{m-1}\Gamma T\nu_{\tau}^{\mathrm{SO}}(n)\to\ldots

And note that here ΩmΓSTντSO(n)\Omega^{m}\Gamma ST\nu_{\tau}^{\mathrm{SO}}(n) equals Ωm1ΓTντSO(n)\Omega^{m-1}\Gamma T\nu_{\tau}^{\mathrm{SO}}(n). This sequence is infinite to the right by construction, but it is also infinite to the left since the number nn could be arbitrary and we get the same maps by applying Ωn+mΓ\Omega^{n+m}\Gamma to the nn’th suspensions as by applying Ωn+m+1Γ\Omega^{n+m+1}\Gamma to the (n+1)(n+1)’st suspensions. We also note that the maps in this sequence commute with the natural maps ΩmΓTντσ(n)ΩmΓTντσ(n+1)\Omega^{m}\Gamma T\nu_{\tau}^{\sigma}(n)\to\Omega^{m}\Gamma T\nu_{\tau}^{\sigma}(n+1) (for σ=SO,int\sigma=\mathrm{SO},\mathrm{int}). Then converging with nn to infinity yields a sequence

ΩmΓTντSOΩmΓTντintΩm1ΓTντSOΩm1ΓTντSO\ldots\to\Omega^{m}\Gamma T\nu_{\tau}^{\mathrm{SO}}\to\Omega^{m}\Gamma T\nu_{\tau}^{\mathrm{int}}\to\Omega^{m-1}\Gamma T\nu_{\tau}^{\mathrm{SO}}\to\Omega^{m-1}\Gamma T\nu_{\tau}^{\mathrm{SO}}\to\ldots

of the direct limits.

If we then fix a manifold QQ and apply the functor [𝑄,][\mathop{\mathop{Q}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\cdot] to this sequence, then we obtain the long exact sequence of cobordism groups as claimed.  \square

Remark 4.2.

In the proof above we used the cofibrations

SnTντSO(n)SnTντint(n)Sn+1TντSO(n)S^{n}T\nu_{\tau}^{\mathrm{SO}}(n)\hookrightarrow S^{n}T\nu_{\tau}^{\mathrm{int}}(n)\to S^{n+1}T\nu_{\tau}^{\mathrm{SO}}(n)

for all nn. Later we will refer to this (and to similar occurences) as a cofibration

TντSOTντintSTντSO.T\nu_{\tau}^{\mathrm{SO}}\hookrightarrow T\nu_{\tau}^{\mathrm{int}}\to ST\nu_{\tau}^{\mathrm{SO}}.

Although these are not actual spaces, applying the functor Γ\Gamma 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 Q=qQ=\mathbb{R}^{q}, then the same long exact sequence could be obtained from the cofibration

TντSOTντintSTντSOT\nu_{\tau}^{\mathrm{SO}}\hookrightarrow T\nu_{\tau}^{\mathrm{int}}\to ST\nu_{\tau}^{\mathrm{SO}}

by just applying the functor Γ\Gamma which yields a fibration

ΓTντintΓTντSOΓSTντSO\Gamma T\nu_{\tau}^{\mathrm{int}}\mathop{\xrightarrow{\leavevmode\nobreak\ \Gamma T\nu_{\tau}^{\mathrm{SO}}\leavevmode\nobreak\ }}\Gamma ST\nu_{\tau}^{\mathrm{SO}}

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 φm\varphi_{m}, χm\chi_{m} and ψm\psi_{m}

First fix a manifold KK, a vector bundle ζK\zeta\to K of rank rr, a closed submanifold AKA\subset K of codimension mm and with normal bundle ξ\xi and put B:=KAB:=K\setminus A. If we denote by Embζ(n,Pn+r)\textstyle{\mathop{\rm Emb}}^{\zeta}(n,P^{n+r}) the cobordism semigroup of nn-manifolds embedded into a fixed manifold Pn+rP^{n+r} with normal bundle induced from ζ\zeta (i.e. ζ\zeta-embeddings), then we have

Embζ(n,P)[𝑃,Tζ].\textstyle{\mathop{\rm Emb}}^{\zeta}(n,P)\cong[\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},T\zeta].

Hence the Puppe sequence of the cofibration Tζ|BTζTζ/Tζ|B=T(ζ|Aξ)T\zeta|_{B}\hookrightarrow T\zeta\to T\zeta/T\zeta|_{B}=T(\zeta|_{A}\oplus\xi) gives us a sequence of homomorphisms

Embζ|B(n,P)Embζ(n,P)Embζ|Aξ(nm,P)Embζ|Bε1(n1,P)\textstyle{\mathop{\rm Emb}}^{\zeta|_{B}}(n,P)\to\textstyle{\mathop{\rm Emb}}^{\zeta}(n,P)\to\textstyle{\mathop{\rm Emb}}^{\zeta|_{A}\oplus\xi}(n-m,P)\mathop{\xrightarrow{\leavevmode\nobreak\ \partial\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Emb}}^{\zeta|_{B}\oplus\varepsilon^{1}}(n-1,P)\to\ldots

of cobordism semigroups (which is exact in the stable dimensions).

Lemma 5.1.

Let f:MnmPn+rf\colon M^{n-m}\hookrightarrow P^{n+r} be a (ζ|Aξ)(\zeta|_{A}\oplus\xi)-embedding which yields a decomposition νf=(ζ|A)MξM\nu_{f}=(\zeta|_{A})_{M}\oplus\xi_{M} (with (ζ|A)M(\zeta|_{A})_{M} and ξM\xi_{M} induced from ζ|A\zeta|_{A} and ξ\xi respectively) and the exponential map can be used to extend ff to an embedding f~\tilde{f} of the total space of ξM\xi_{M} into a small neighbourhood of f(M)f(M). Then the boundary homomorphism \partial assigns to the cobordism class [f][f] the class of f~|SξM\tilde{f}|_{S\xi_{M}} which has codimension r+1r+1, 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 ζ|B\zeta|_{B}.

Proof.  We obtain T(ζ|Aξ)T(\zeta|_{A}\oplus\xi) in the Puppe sequence by restricting ζ\zeta to a tubular neighbourhood UKU\subset K of AA and then factoring TζT\zeta by the complement of this restriction, that is, attaching the cone CT(ζ|KU)CT(\zeta|_{K\setminus U}). The next item in the sequence is ST(ζ|B)ST(\zeta|_{B}) which we get by gluing the cone CTζCT\zeta to the previously obtained space. In the following we shall always mean the above representations of the spaces T(ζ|Aξ)T(\zeta|_{A}\oplus\xi) and ST(ζ|B)ST(\zeta|_{B}).

Refer to caption
Figure 1: We show the local form of T(ζ|Aξ)T(\zeta|_{A}\oplus\xi) (on the left) and ST(ζ|B)ST(\zeta|_{B}) (on the right) in the case r=0r=0, m=1m=1.

If the homotopy class of κf:𝑃T(ζ|Aξ)\kappa_{f}\colon\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\to T(\zeta|_{A}\oplus\xi) induces the cobordism class of f:MPf\colon M\hookrightarrow P (such that f(M)=κf1(A)f(M)=\kappa_{f}^{-1}(A)), then its image under

:[𝑃,T(ζ|Aξ)][𝑃,ST(ζ|B)]\partial\colon[\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},T(\zeta|_{A}\oplus\xi)]\to[\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},ST(\zeta|_{B})]

is represented by λf:𝑃ST(ζ|B)\lambda_{f}\colon\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\to ST(\zeta|_{B}) which we get by taking the portion of the image of κf\kappa_{f} in the total space of ζ|U\zeta|_{U} and pushing it toward the special point of the cone CT(ζ|U)CTζCT(\zeta|_{U})\subset CT\zeta, in this way detaching it from AA, and otherwise leaving the rest of κf\kappa_{f} unchanged.

Refer to caption
Figure 2: We show the local form of the image of κf\kappa_{f} (on the left) and λf\lambda_{f} (on the right) in ST(ζ|B)ST(\zeta|_{B}) again in the case r=0r=0, m=1m=1.

The inducing map κf\kappa_{f} is such that the exponential image of the normal bundle νf=(ζ|A)MξM\nu_{f}=(\zeta|_{A})_{M}\oplus\xi_{M} (restricted to a small neighbourhood) is the preimage of a tubular neighbourhood of AT(ζ|Aξ)A\subset T(\zeta|_{A}\oplus\xi) and it is mapped fibrewise bijectively to its image in this neighbourhood. Thus we may assume that the image of f~\tilde{f} coincides with κf1(U)\kappa_{f}^{-1}(U) such that f~(SξM)=κf1(U)\tilde{f}(S\xi_{M})=\kappa_{f}^{-1}(\partial U) and κf\kappa_{f} is again a fibrewise bijection from SξMS\xi_{M} to U|κf(κf1(A))\partial U|_{\kappa_{f}(\kappa_{f}^{-1}(A))}.

Then we can assume that the image of λf\lambda_{f} intersects the base space KK only in U\partial U (this intersection is in B=KAB=K\setminus A) and λf\lambda_{f} restricted to the preimage of this intersection coincides with κf\kappa_{f}. Hence for any point p(κff)1(A)p\in(\kappa_{f}\circ f)^{-1}(A) we have that λf\lambda_{f} maps the sphere f~(SpξM)\tilde{f}(S_{p}\xi_{M}) to BB and the union of these (that is, f~(SξM)\tilde{f}(S\xi_{M})) is the whole λf1(B)\lambda_{f}^{-1}(B). Since SξMS\xi_{M} 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 ST(ζ|B)ST(\zeta|_{B}). The orthogonal complement of this in the normal bundle of f~(SξM)\tilde{f}(S\xi_{M}) is then induced from the ζ|B\zeta|_{B} part of the normal bundle of BST(ζ|B)B\subset ST(\zeta|_{B}). This is what we wanted to prove.  \square

If the manifold above has the form K=S1×2VK=S^{1}\mathop{\mathop{\times}\limits_{\vbox to-1.50694pt{\kern-5.425pt\hbox{$\scriptstyle\mathbb{Z}_{2}$}\vss}}}V and the submanifold AA is a fibre VV, then we have ξ=ε1\xi=\varepsilon^{1}, BVB\cong V and ζ|Bζ|A=ζ|V\zeta|_{B}\cong\zeta|_{A}=\zeta|_{V}, hence the sequence above has the form

Embζ|V(n,P)Embζ(n,P)Embζ|Vε1(n1,P)Embζ|Vε1(n1,P)\textstyle{\mathop{\rm Emb}}^{\zeta|_{V}}(n,P)\to\textstyle{\mathop{\rm Emb}}^{\zeta}(n,P)\to\textstyle{\mathop{\rm Emb}}^{\zeta|_{V}\oplus\varepsilon^{1}}(n-1,P)\mathop{\xrightarrow{\leavevmode\nobreak\ \partial\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Emb}}^{\zeta|_{V}\oplus\varepsilon^{1}}(n-1,P)\to\ldots
Corollary 5.2.

The boundary homomorphism \partial in this case is id+ι\mathop{\rm id}+\iota where ι\iota is the involution which acts on a cobordism class represented by a 11-framed ζ|V\zeta|_{V}-embedding f:Mn1Pn+rf\colon M^{n-1}\hookrightarrow P^{n+r} such that it inverts its framing vector field vv and in the orthogonal complement of vv in νf\nu_{f} (induced from ζ|V\zeta|_{V}) it applies the 2\mathbb{Z}_{2}-action of VV.

Proof.  We want to understand the embedding f~|SξM\tilde{f}|_{S\xi_{M}} in the lemma above in this special case. Now ξ\xi is a trivial line bundle, so SζM=MM+S\zeta_{M}=M_{-}\sqcup M_{+} where the M±M_{\pm} are both diffeomorphic to MM, the maps f±:=f~|M±f_{\pm}:=\tilde{f}|_{M_{\pm}} can be identified with ff and the normal vector fields v±v_{\pm} giving the 11-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 11-framing v+v_{+} of f+f_{+} is the same as the 11-framing vv of ff and vv_{-} is opposite to it.

What is left is to understand how the orthogonal complement of v±v_{\pm} in the normal bundle is induced from ζ|V\zeta|_{V} on the two manifolds M±M_{\pm}. If our fixed fibre was A={p}×VA=\{p\}\times V for a point pS1p\in S^{1}, then (with the notation of the previous proof) the tubular neighbourhood UKU\subset K is (pε,p+ε)×V(p-\varepsilon,p+\varepsilon)\times V for a small number ε\varepsilon. Now f(M)=κf1({p}×V)f(M)=\kappa_{f}^{-1}(\{p\}\times V) and the part of νf\nu_{f} induced from ζ|V\zeta|_{V} is pulled back by the restriction of κf\kappa_{f} from the bundle over {p}×V\{p\}\times V. Identifying the fibres in [pε,p+ε]×V[p-\varepsilon,p+\varepsilon]\times V in the trivial way then yields that the ζ|V\zeta|_{V} parts of the normal bundles of f±(M±)=κf1({p±ε}×V)f_{\pm}(M_{\pm})=\kappa_{f}^{-1}(\{p\pm\varepsilon\}\times V) are induced by the same map from the bundle over {p±ε}×V\{p\pm\varepsilon\}\times V. But then identifying the fibres over S1{p}S^{1}\setminus\{p\} (i.e. in BB) instead of [pε,p+ε][p-\varepsilon,p+\varepsilon] changes the identification of {p+ε}×V\{p+\varepsilon\}\times V with {pε}×V\{p-\varepsilon\}\times V by applying the 2\mathbb{Z}_{2}-action in one of them, say in {pε}×V\{p-\varepsilon\}\times V.

Thus we got that both the 11-framing of f+f_{+} and the inducing map of the rest of its normal bundle coincide with those of ff, on the other hand, the 11-framing of ff_{-} is opposite to that of ff and the inducing map of the rest of its normal bundle is composed with the 2\mathbb{Z}_{2}-action in VV. Hence we have [(f,v)]=[(f+,v+)]+[(f,v)]\partial[(f,v)]=[(f_{+},v_{+})]+[(f_{-},v_{-})] where [(f+,v+)]=[(f,v)][(f_{+},v_{+})]=[(f,v)] and [(f,v)]=ι[(f,v)][(f_{-},v_{-})]=\iota[(f,v)] for the involution ι\iota described in our statement above.  \square

Proposition 5.3.

φm:CobτSO(Q×m)CobτSO(Q×m)\varphi_{m}\colon\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m})\to\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m}) is of the form id+ι\mathop{\rm id}+\iota where ι\iota is the involution which, when representing cobordism classes by τSO\tau^{\mathrm{SO}}-embeddings (see remark 3.7), composes a map with the reflection to a hypersurface Q×m+r1Q×m+rQ\times\mathbb{R}^{m+r-1}\subset Q\times\mathbb{R}^{m+r} but does not change its orientation.

Proof.  Let qq be the dimension of QQ and put n:=q+mkn:=q+m-k. Then remark 3.7 and theorem 3.11 show that for σ=int,SO\sigma=\mathrm{int},\mathrm{SO} the group Cobτσ(Qq×m)\textstyle{\mathop{\rm Cob}}_{\tau}^{\sigma}(Q^{q}\times\mathbb{R}^{m}) is naturally isomorphic to

Embτσ(n,Q×m+r)Embντσ(r)εr(n,Q×m+r)\textstyle{\mathop{\rm Emb}}_{\tau}^{\sigma}(n,Q\times\mathbb{R}^{m+r})\cong\textstyle{\mathop{\rm Emb}}^{\nu_{\tau}^{\sigma}(r)\oplus\varepsilon^{r}}(n,Q\times\mathbb{R}^{m+r})

where rr is a sufficiently large number. Observe also that we have Kτint=S1×2KτSOK_{\tau}^{\mathrm{int}}=S^{1}\mathop{\mathop{\times}\limits_{\vbox to-1.50694pt{\kern-5.425pt\hbox{$\scriptstyle\mathbb{Z}_{2}$}\vss}}}K_{\tau}^{\mathrm{SO}} and, using the notation of the proof of theorem 4.1, we let K1K_{1} be the fibre of KτintK_{\tau}^{\mathrm{int}} over a point in S1S^{1} and K2K_{2} is its complement in KτintK_{\tau}^{\mathrm{int}} and for i=1,2i=1,2 we let the virtual bundle νi\nu_{i} be the restriction of ντint\nu_{\tau}^{\mathrm{int}} over KiK_{i} (hence KiKτSOK_{i}\cong K_{\tau}^{\mathrm{SO}} and νiντSO\nu_{i}\cong\nu_{\tau}^{\mathrm{SO}}). We now think of these Kazarian spaces as finite dimensional approximations of the actual spaces (hence closed manifolds) over which the (r1)(r-1)’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 φm\varphi_{m} are

[Sm+r𝑄,SrTν1]and[Sm+r𝑄,SrTν2][S^{m+r}\mathop{\mathop{Q}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T\nu_{1}]\quad\text{and}\quad[S^{m+r}\mathop{\mathop{Q}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T\nu_{2}]

respectively, thus we are in the setting of corollary 5.2 with K=KτintK=K_{\tau}^{\mathrm{int}}, V=KτSOV=K_{\tau}^{\mathrm{SO}}, A=K1A=K_{1}, B=K2B=K_{2} and ζ=ντintεr\zeta=\nu_{\tau}^{\mathrm{int}}\oplus\varepsilon^{r} with ζ|V=ντSOεr\zeta|_{V}=\nu_{\tau}^{\mathrm{SO}}\oplus\varepsilon^{r}, and so the boundary homomorphism =φm\partial=\varphi_{m} is indeed id+ι\mathop{\rm id}+\iota. The involution ι\iota here inverts one (say, the last) of the framing vectors of any τSO\tau^{\mathrm{SO}}-embedding into Q×m+rQ\times\mathbb{R}^{m+r} (see remark 3.7) and applies the 2\mathbb{Z}_{2}-action of KτSOK_{\tau}^{\mathrm{SO}} in the orthogonal complement of this vector in the normal bundle.

But if rr is large enough, then we may assume for any τSO\tau^{\mathrm{SO}}-embedding to map into a hypesurface Q×r+m1Q×r+mQ\times\mathbb{R}^{r+m-1}\subset Q\times\mathbb{R}^{r+m} 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 2\mathbb{Z}_{2}-action of KτSOK_{\tau}^{\mathrm{SO}} reverses the orientation in the orthogonal complement too, which means that the orientation of the normal bundle remains unchanged. This finishes the proof.  \square

Proposition 5.4.

χm:CobτSO(Q×m)Cobτint(Q×m)\chi_{m}\colon\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m})\to\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m}) is the forgetful homomorphism that assigns to an oriented cobordism class [f][f] the class of (f,0)(f,0) as a Wall map.

Proof.  This follows immediately since the map between the classifying spaces is just the inclusion of ΓTν2\Gamma T\nu_{2} into ΓTντint\Gamma T\nu_{\tau}^{\mathrm{int}} (see the proof of theorem 4.1).  \square

Proposition 5.5.

ψm:Cobτint(Q×m)CobτSO(Q×m1)\psi_{m}\colon\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m})\to\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m-1}) assigns to a cobordism class [(f,w)][(f,w)] the oriented cobordism class of ff restricted to the Poincaré dual of ww which can be represented by a 11-codimensional submanifold uniquely up to cobordism, the restriction f|PD(w)f|_{\mathop{\rm PD}(w)} can be assumed to be a mapping to Q×m1Q\times\mathbb{R}^{m-1}, it has oriented normal bundle and its cobordism class only depends on the class of (f,w)(f,w), hence this assignment is well-defined.

Proof.  The cobordism class [(f,w)][(f,w)] is represented by a τ\tau-map f:MQ×mf\colon M\to Q\times\mathbb{R}^{m} together with a fixed cohomology class wH1(M;)w\in H^{1}(M;\mathbb{Z}) such that its mod 22 reduction is w1(νf)w_{1}(\nu_{f}). Then we have a commutative diagram

S1\textstyle{S^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w\scriptstyle{w}w1(νf)\scriptstyle{w_{1}(\nu_{f})}P\textstyle{\mathbb{R}P^{\infty}}

The Poincaré dual of ww is represented by the codimension-11 submanifold N:=w1(p)MN:=w^{-1}(p)\subset M for a point pS1p\in S^{1}. Note that the normal bundle of NN in MM is trivial and so f|Nf|_{N} has a normal vector field which is uniquely defined by a fixed normal vector of pp in S1S^{1}; we also have w1(νf|N)=w1(νf)|N=0w_{1}(\nu_{f|_{N}})=w_{1}(\nu_{f})|_{N}=0, hence f|Nf|_{N} is orientable and can be canonically oriented (since the classifying map of the stabilisation of νf\nu_{f} maps to BintB\mathrm{int}, hence that of νf|N\nu_{f|_{N}} maps to the fibre B1BSOB_{1}\cong B\mathrm{SO} of BintB\mathrm{int} over pS1p\in S^{1}); finally we may assume that NN intersects each singularity stratum in MM transversally, thus f|Nf|_{N} is a τSO1\tau^{\mathrm{SO}\oplus 1}-map.

By (2.1) we have CobτSO1(Q×m)CobτSO(Q×m1)\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}\oplus 1}(Q\times\mathbb{R}^{m})\cong\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m-1}) which means that assigning to the cobordism class of (f,w)(f,w) the class of f|Nf|_{N} gives a map

Cobτint(Q×m)CobτSO(Q×m1)\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m})\to\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{m-1})

taking [f][f|PD(w)][f]\mapsto[f|_{\mathop{\rm PD}(w)}]. We now only need to prove that this map is well-defined and is the same as ψm\psi_{m} for which it is sufficient to prove that ψm\psi_{m} assigns to the classifying map of ff the classifying map of f|Nf|_{N}.

We again refer to remark 3.7: [f][f] bijectively corresponds to a homotopy class [κf][Sm𝑄,ΓTντint][\kappa_{f}]\in[S^{m}\mathop{\mathop{Q}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Gamma T\nu_{\tau}^{\mathrm{int}}] represented by κf:Sr+m𝑄SrTντint\kappa_{f}\colon S^{r+m}\mathop{\mathop{Q}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\to S^{r}T\nu_{\tau}^{\mathrm{int}} where rr is a large number and we again consider such a finite dimensional approximation of the Kazarian space over which SrTντintS^{r}T\nu_{\tau}^{\mathrm{int}} exists. In the proof of theorem 4.1 we obtained the map in the Puppe sequence inducing ψm\psi_{m} by factoring T(ντintεr)T(\nu_{\tau}^{\mathrm{int}}\oplus\varepsilon^{r}) by its restriction over the complement of a tubular neighbourhood of K1KτintK_{1}\subset K_{\tau}^{\mathrm{int}}. Recall that K1KτSOK_{1}\cong K_{\tau}^{\mathrm{SO}} is the preimage of B1BSOBintB_{1}\cong B\mathrm{SO}\subset B\mathrm{int} (the fibre of BintB\mathrm{int} over pS1p\in S^{1}). Thus we have a diagram

T(ντintεr)\textstyle{T(\nu_{\tau}^{\mathrm{int}}\oplus\varepsilon^{r})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}Tγr+kint\textstyle{T\gamma^{\mathrm{int}}_{r+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ST(ν1εr)\textstyle{ST(\nu_{1}\oplus\varepsilon^{r})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}STγr+kSO\textstyle{ST\gamma^{\mathrm{SO}}_{r+k}}Kτint\textstyle{K_{\tau}^{\mathrm{int}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bint(r+k)\textstyle{B\mathrm{int}(r+k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S1\textstyle{S^{1}}K1\textstyle{K_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B1(r+k)\textstyle{B_{1}(r+k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\textstyle{p\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces} \in

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 ψm\psi_{m} assigns to the homotopy class of the inducing map κf\kappa_{f} the homotopy class of its composition with the quotient map qq. Now κf\kappa_{f} pulls back the τint\tau^{\mathrm{int}}-embedding f×if\times i (for an embedding i:Mri\colon M\hookrightarrow\mathbb{R}^{r}) from the embedding of KτintK_{\tau}^{\mathrm{int}} into T(ντintεr)T(\nu_{\tau}^{\mathrm{int}}\oplus\varepsilon^{r}) and the composition qκfq\circ\kappa_{f} pulls back a representative of ψm[f×i]\psi_{m}[f\times i] from the embedding of K1K_{1} into ST(ν1εr)ST(\nu_{1}\oplus\varepsilon^{r}). But note that we have N=(qκf)1(K1)N=(q\circ\kappa_{f})^{-1}(K_{1}) and its τSO\tau^{\mathrm{SO}}-embedding into Q×r+mQ\times\mathbb{R}^{r+m} is f|N×if|_{N}\times i. Hence we proved that ψm[f]=[f|N]\psi_{m}[f]=[f|_{N}] and this is what we needed.  \square

This finishes the proof of theorem 2.

Remark 5.6.

If the codimension kk is sufficiently large (compared to nn), then we have CobτSO(n+k)=Ωn\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(\mathbb{R}^{n+k})=\Omega_{n} and Cobτint(n+k)=𝔚n\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(\mathbb{R}^{n+k})=\mathfrak{W}_{n} for any (non-empty) singularity set τ\tau. Hence in the case Q×m=n+kQ\times\mathbb{R}^{m}=\mathbb{R}^{n+k} the portion of the long exact sequence in theorem 2 where nn is sufficiently small (compared to kk) looks like

ΩnΩn𝔚nΩn1\ldots\to\Omega_{n}\to\Omega_{n}\to\mathfrak{W}_{n}\to\Omega_{n-1}\to\ldots

If we fix here nn, then kk 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).

Remark 5.7.

It is easy to see that propositions 5.3 and 5.5 also extend [li, lemma 2].

Part II The unoriented case

6 Cobordism of τGξ\tau^{G\oplus\xi}-maps

Fix a stable group GG which is the direct limit of the groups G(n)G(n). Recall that G(n)G(n) has a fixed linear action on n\mathbb{R}^{n} and γnGBG(n)\gamma_{n}^{G}\to BG(n) denotes the universal rank-nn vector bundle with structure group GG. Fix also a rank-mm vector bundle ξ\xi over the space BG=limnBG(n)BG=\mathop{\lim}\limits_{\vbox to-1.50694pt{\kern-4.82224pt\hbox{$\scriptstyle n\to\infty$}\vss}}BG(n).

Definition 6.1.

Let τ\tau be a set of stable singularities. We define a τGξ\tau^{G\oplus\xi}-map from a compact manifold MnM^{n} to a manifold Pn+k+mP^{n+k+m} to be the germ along the zero-section of a τG\tau^{G}-map

f~:ξMP\tilde{f}\colon\xi_{M}\to P

where ξM\xi_{M} is a rank-mm vector bundle over MM and f~\tilde{f} has the following properties:

  • (i)

    the differential df~d\tilde{f} restricted to any fibre of ξM\xi_{M} is injective,

  • (ii)

    noting that (i) implies ν(f~|M)=(νf~)|MξM\nu_{(\tilde{f}|_{M})}=(\nu_{\tilde{f}})|_{M}\oplus\xi_{M} we require that the map MBGM\to BG inducing the (virtual) GG-bundle (νf~)|M(\nu_{\tilde{f}})|_{M} also pulls back ξM\xi_{M} from ξ\xi, thus it pulls back ν(f~|M)εr\nu_{(\tilde{f}|_{M})}\oplus\varepsilon^{r} from γk+rGξ\gamma^{G}_{k+r}\oplus\xi for sufficiently large numbers rr.

The cobordism of two τGξ\tau^{G\oplus\xi}-maps is defined in the usual way (i.e. as a τGξ\tau^{G\oplus\xi}-map to P×[0,1]P\times[0,1] where both the germ and the map to BGBG inducing the vector bundle extend those of the boundary) and the cobordism group of τGξ\tau^{G\oplus\xi}-maps to the manifold PP is denoted by CobτGξ(P)\textstyle{\mathop{\rm Cob}}_{\tau}^{G\oplus\xi}(P).

Example 6.2.

In the case G=SOG=\mathrm{SO} and ξ=εm\xi=\varepsilon^{m}, the above definition is exactly the definition of mm-framed τ\tau-maps in [hosszu].

As we saw previously in part I, the main tools for understanding how cobordisms of τ\tau-maps are induced from the classifying spaces are τ\tau-embeddings. In order to construct the classifying space of τGξ\tau^{G\oplus\xi}-maps we now introduce the following:

Definition 6.3.

Let τ\tau be a set of stable singularities. By a τGξ\tau^{G\oplus\xi}-embedding of a manifold MnM^{n} into a manifold QqQ^{q} we mean a quadruple (e,𝒱,ξM,)(e,\mathscr{V},\xi_{M},\mathscr{F}) where

  • (i)

    e:MQe\colon M\hookrightarrow Q is an embedding,

  • (ii)

    𝒱=(v1,,vr)\mathscr{V}=(v_{1},\ldots,v_{r}) where r=qnkmr=q-n-k-m and the viv_{i}’s are pointwise independent vector fields along e(M)e(M), i.e. sections of the bundle TQ|e(M)TQ|_{e(M)}; we identify 𝒱\mathscr{V} with the trivialised subbundle generated by the viv_{i}’s,

  • (iii)

    ξM\xi_{M} is a rank-mm subbundle of TQ|e(M)TQ|_{e(M)} which is pointwise independent of both 𝒱\mathscr{V} and Te(M)Te(M) and the normal bundle νe=νξM\nu_{e}=\nu^{\prime}\oplus\xi_{M} is induced from the bundle γk+rGξ|BG(k+r)\gamma_{k+r}^{G}\oplus\xi|_{BG(k+r)} over BG(k+r)BG(k+r) by a map MBG(k+r)M\to BG(k+r) which pulls back ξM\xi_{M} from ξ|BG(k+r)\xi|_{BG(k+r)} and ν\nu^{\prime} from γk+rG\gamma_{k+r}^{G},

  • (iv)

    \mathscr{F} is a foliation of dimension r+mr+m on a neighbourhood of e(M)e(M) and it is tangent to 𝒱ξM\mathscr{V}\oplus\xi_{M} along e(M)e(M),

  • (v)

    any point pMp\in M has a neighbourhood on which the composition of ee with the projection along the leaves of \mathscr{F} to a small (n+k)(n+k)-dimensional transverse slice has at pp a singularity which belongs to τ\tau.

The cobordism of two τGξ\tau^{G\oplus\xi}-embeddings is defined in the usual way and the cobordism group of τGξ\tau^{G\oplus\xi}-embeddings of nn-manifolds to QQ is denoted by EmbτGξ(n,Q)\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,Q).

Remark 6.4.

Such a τGξ\tau^{G\oplus\xi}-embedding induces a stratification of MM by the submanifolds

η(e):=η(e,𝒱,ξM,):={pMpη(πe)}\eta(e):=\eta(e,\mathscr{V},\xi_{M},\mathscr{F}):=\{p\in M\mid p\in\eta(\pi\circ e)\}

where the [η][\eta] are the elements of τ\tau and π\pi denotes the local projection around e(p)e(p) along the leaves of \mathscr{F}.

Example 6.5.

If ξM\xi_{M} is a vector bundle over MnM^{n}, f~:ξMPn+k+m\tilde{f}\colon\xi_{M}\to P^{n+k+m} is a τGξ\tau^{G\oplus\xi}-map and i:Mri\colon M\hookrightarrow\mathbb{R}^{r} is an embedding, then we can define a τGξ\tau^{G\oplus\xi}-embedding (e,𝒱,ξM,)(e,\mathscr{V},\xi_{M},\mathscr{F}) of MM into P×rP\times\mathbb{R}^{r}: We choose an arbitrarily small representative ff of f~\tilde{f} and put e:=f|M×ie:=f|_{M}\times i; the vector fields viv_{i} arise from a basis in r\mathbb{R}^{r}; the bundle ξM\xi_{M} can now be viewed as a subbundle of T(P×r)|e(M)T(P\times\mathbb{R}^{r})|_{e(M)} (since for any pMp\in M we can identify ξM|p\xi_{M}|_{p} with the translate of dfp(ξM|p)df_{p}(\xi_{M}|_{p}) to Te(p)(P×r)T_{e(p)}(P\times\mathbb{R}^{r})); and \mathscr{F} is composed of the leaves f(ξM|p)×rf(\xi_{M}|_{p})\times\mathbb{R}^{r} intersected by a small neighbourhood of e(M)e(M) (for all pMp\in M).

Definition 6.6.
  1. (1)

    The vector fields 𝒱=(v1,,vr)\mathscr{V}=(v_{1},\ldots,v_{r}) and the foliation \mathscr{F} in the above example are called vertical in P×rP\times\mathbb{R}^{r}.

  2. (2)

    The subsets P×{x}P×rP\times\{x\}\subset P\times\mathbb{R}^{r} (for any xrx\in\mathbb{R}^{r}) are called horizontal sections.

The rest of this section consists of constructing the classifying space of the cobordism group of τGξ\tau^{G\oplus\xi}-maps in a similar way as sketched in remark 3.7. First we state a lemma and a theorem which connect τGξ\tau^{G\oplus\xi}-maps with τGξ\tau^{G\oplus\xi}-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 τ\tau be a set of stable singularities and (e,𝒱,ξM,)(e,\mathscr{V},\xi_{M},\mathscr{F}) be a τGξ\tau^{G\oplus\xi}-embedding of MnM^{n} into Pn+k+m×rP^{n+k+m}\times\mathbb{R}^{r} where MM is a compact manifold and PP is any manifold. Then there is a diffeotopy Φt(t[0,1])\Phi_{t}\leavevmode\nobreak\ (t\in[0,1]) of P×rP\times\mathbb{R}^{r} such that Φ0\Phi_{0} is the identity and (the differential of) Φ1\Phi_{1} takes 𝒱\mathscr{V} to the vertical vector fields 𝒱\mathscr{V}^{\prime} and \mathscr{F} to the vertical foliation \mathscr{F}^{\prime} around the image of MM. The relative version of this claim is also true, that is, if the vector fields 𝒱\mathscr{V} and the foliation \mathscr{F} are already vertical on a neighbourhood of a compact subset Ce(M)C\subset e(M), then the diffeotopy Φt(t[0,1])\Phi_{t}\leavevmode\nobreak\ (t\in[0,1]) is fixed on a neighbourhood of CC.

Proof.  The manifold MM is finitely stratified by the submanifolds

Si:=[η]τdimη(e)=iη(e),i=0,,n.S_{i}:=\bigcup_{[\eta]\in\tau\atop\dim\eta(e)=i}\eta(e),\quad i=0,\ldots,n.

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 P×rP\times\mathbb{R}^{r} which turns the vector fields 𝒱\mathscr{V} into vertical vector fields. Therefore we may assume that 𝒱\mathscr{V} is already vertical and so we can also assume that the fibres of ξM\xi_{M} 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 \mathscr{F} into the vertical foliation \mathscr{F}^{\prime} and its differential keeps the vector fields 𝒱=𝒱\mathscr{V}=\mathscr{V}^{\prime} vertical.

We will recursively deform \mathscr{F} into \mathscr{F}^{\prime} around the images of the strata Si(i=0,,n)S_{i}\leavevmode\nobreak\ (i=0,\ldots,n). First we list some general observations:

  1. (1)

    If RrR\subset\mathbb{R}^{r} and L,LP×RL,L^{\prime}\subset P\times R are such that each of LL and LL^{\prime} intersects each horizontal section P×{x}(xR)P\times\{x\}\leavevmode\nobreak\ (x\in R) exactly once, then a bijective correspondence LLL\to L^{\prime} arises by associating the points on the same horizontal section to each other.

  2. (2)

    If AP×rA\subset P\times\mathbb{R}^{r} is such that for each aAa\in A subsets Ra,La,LaR_{a},L_{a},L^{\prime}_{a} are given as in (1), then a family of bijective maps {LaLaaA}\{L_{a}\to L^{\prime}_{a}\mid a\in A\} arises. If we have La1La2==La1La2L_{a_{1}}\cap L_{a_{2}}=\varnothing=L^{\prime}_{a_{1}}\cap L^{\prime}_{a_{2}} for any two different points a1,a2Aa_{1},a_{2}\in A, then the union of these bijections gives a continuous bijective map

    φ:U:=aALaaALa=:U\varphi\colon U:=\bigcup_{a\in A}L_{a}\to\bigcup_{a\in A}L^{\prime}_{a}=:U^{\prime}
  3. (3)

    If the subsets A,La,LaA,L_{a},L^{\prime}_{a} in (2) are submanifolds of P×rP\times\mathbb{R}^{r} such that UU and UU^{\prime} are also submanifolds, then the map φ\varphi is smooth. In this case for all points (p,x)U(p,x)\in U we can join (p,x)(p,x) and φ(p,x)\varphi(p,x) by a minimal geodesic in the horizontal section P×{x}P\times\{x\}, and using these we can extend φ\varphi to an isotopy φt(t[0,1])\varphi_{t}\leavevmode\nobreak\ (t\in[0,1]) of UU (for which φ0=idU\varphi_{0}=\mathop{\rm id}_{U} and φ1=φ\varphi_{1}=\varphi).

Denote by 𝒱\mathscr{V}^{\perp} the orthogonal complement of the bundle 𝒱|e(S0)Te(S0)\mathscr{V}|_{e(S_{0})}\oplus Te(S_{0}) in T(P×r)|e(S0)T(P\times\mathbb{R}^{r})|_{e(S_{0})} (with respect to some Riemannian metric). Choose a small neighbourhood AA of e(S0)e(S_{0}) in exp(𝒱)\exp(\mathscr{V}^{\perp}) (where exp\exp denotes the exponential map of P×rP\times\mathbb{R}^{r}) and for all aAa\in A let LaL_{a} and LaL^{\prime}_{a} be the intersections of a small neighbourhood of aa and the leaves of \mathscr{F} and \mathscr{F}^{\prime} respectively.

If the neighbourhoods were chosen sufficiently small, then we are in the setting of (3), hence a diffeomorphism φ:UU\varphi\colon U\to U^{\prime} arises (with the same notations as above). Note that UU and UU^{\prime} are both neighbourhoods of e(S0)e(S_{0}), the map φ\varphi fixes e(S0)e(S_{0}) and for all ae(S0)a\in e(S_{0}) we have dφa=idTa(P×m)d\varphi_{a}=\mathop{\rm id}_{T_{a}(P\times\mathbb{R}^{m})}. Observe that where the foliations \mathscr{F} and \mathscr{F}^{\prime} initially coincide, this method just gives the identity for all t[0,1]t\in[0,1]. The isotopy we get this way can be extended to a diffeotopy of P×rP\times\mathbb{R}^{r} (by the isotopy extension theorem) and it takes the leaves of \mathscr{F} to the leaves of \mathscr{F}^{\prime} around the image of S0S_{0}.

Next we repeat the same procedure around e(S1)e(S_{1}), the image of the next stratum, to get a new diffeotopy (that leaves a neighbourhood of e(S0)e(S_{0}) unchanged), and so on. In the end we obtain a diffeotopy of P×rP\times\mathbb{R}^{r} which turns \mathscr{F} into the vertical foliation \mathscr{F}^{\prime} around the image of MM and does not change the vertical vector fields 𝒱=𝒱\mathscr{V}=\mathscr{V}^{\prime}.  \square

Theorem 6.8.

For any set τ\tau of stable singularities and any manifold Pn+k+mP^{n+k+m}, if the number rr is sufficiently large (compared to nn), then we have

CobτGξ(P)EmbτGξ(n,P×r).\textstyle{\mathop{\rm Cob}}_{\tau}^{G\oplus\xi}(P)\cong\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r}).

Proof.  Take any number r2n+4r\geq 2n+4, so any manifold of dimension at most n+1n+1 can be embedded into r\mathbb{R}^{r} uniquely up to isotopy. We will define two homomorphisms α:CobτGξ(P)EmbτGξ(n,P×r)\alpha\colon\textstyle{\mathop{\rm Cob}}_{\tau}^{G\oplus\xi}(P)\to\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r}) and β:EmbτGξ(n,P×r)CobτGξ(P)\beta\colon\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r})\to\textstyle{\mathop{\rm Cob}}_{\tau}^{G\oplus\xi}(P) which will turn out to be each other’s inverses.

I. Construction of α:CobτGξ(P)EmbτGξ(n,P×r)\alpha\colon\textstyle{\mathop{\rm Cob}}_{\tau}^{G\oplus\xi}(P)\to\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r}).

For a τGξ\tau^{G\oplus\xi}-map f~:ξMPn+k\tilde{f}\colon\xi_{M}\to P^{n+k} we can choose any embedding i:Mnri\colon M^{n}\hookrightarrow\mathbb{R}^{r} and define a τGξ\tau^{G\oplus\xi}-embedding (e,𝒱,ξM,)(e,\mathscr{V},\xi_{M},\mathscr{F}) as in example 6.5. Define the map α\alpha to assign to the cobordism class of ff the cobordism class of e=(e,𝒱,ξM,)e=(e,\mathscr{V},\xi_{M},\mathscr{F}). In order to prove that α\alpha is well-defined, we have to show that the cobordism class of ee does not depend on the choice of the embedding ii and the representative of the class [f][f].

Claim.  If i0:Mri_{0}\colon M\hookrightarrow\mathbb{R}^{r} and i1:Mri_{1}\colon M\hookrightarrow\mathbb{R}^{r} are two embeddings and the above method assigns to them the τGξ\tau^{G\oplus\xi}-embeddings e0=(e0,𝒱0,ξM,0)e_{0}=(e_{0},\mathscr{V}_{0},\xi_{M},\mathscr{F}_{0}) and e1=(e1,𝒱1,ξM,1)e_{1}=(e_{1},\mathscr{V}_{1},\xi_{M},\mathscr{F}_{1}) respectively, then e0e1e_{0}\sim e_{1}.

Proof.  Because of the dimension condition, i0i_{0} and i1i_{1} can be connected by an isotopy it(t[0,1])i_{t}\leavevmode\nobreak\ (t\in[0,1]). We define a τGξ\tau^{G\oplus\xi}-embedding

E:M×[0,1]P×r×[0,1];(p,t)(f(p),it(p),t)E\colon M\times[0,1]\hookrightarrow P\times\mathbb{R}^{r}\times[0,1];\leavevmode\nobreak\ (p,t)\mapsto(f(p),i_{t}(p),t)

(again with the horizontal ξM\xi_{M} and the vertical vector fields and foliation), which is precisely a cobordism between e0e_{0} and e1e_{1}.  \diamond

Claim.  If f~0:ξM0P\tilde{f}_{0}\colon\xi_{M_{0}}\to P and f~1:ξM1P\tilde{f}_{1}\colon\xi_{M_{1}}\to P are cobordant τGξ\tau^{G\oplus\xi}-maps and the above method assigns to them the τGξ\tau^{G\oplus\xi}-embeddings e0=(e0,𝒱0,ξM0,0)e_{0}=(e_{0},\mathscr{V}_{0},\xi_{M_{0}},\mathscr{F}_{0}) and e1=(e1,𝒱1,e_{1}=(e_{1},\mathscr{V}_{1}, ξM1,1)\xi_{M_{1}},\mathscr{F}_{1}) respectively, then e0e1e_{0}\sim e_{1}.

Proof.  Let F~:ξWP×[0,1]\tilde{F}\colon\xi_{W}\to P\times[0,1] be a cobordism between f~0\tilde{f}_{0} and f~1\tilde{f}_{1} (where ξW\xi_{W} is a vector bundle over a manifold Wn+1W^{n+1}). Again by the dimension condition, the embedding i0i1:M0M1=Wri_{0}\sqcup i_{1}\colon M_{0}\sqcup M_{1}=\partial W\hookrightarrow\mathbb{R}^{r} extends to an embedding I:WrI\colon W\hookrightarrow\mathbb{R}^{r}. Hence the map E:=F×IE:=F\times I is a τGξ\tau^{G\oplus\xi}-embedding of WW into P×r×[0,1]P\times\mathbb{R}^{r}\times[0,1] (the vector fields and foliation are again vertical and ξW\xi_{W} is horizontal) and it is easy to see that this is a cobordism between e0e_{0} and e1e_{1}.  \diamond

II. Construction of β:EmbτGξ(n,P×r)CobτGξ(P)\beta\colon\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r})\to\textstyle{\mathop{\rm Cob}}_{\tau}^{G\oplus\xi}(P).

If e=(e,𝒱,ξM,)e=(e,\mathscr{V},\xi_{M},\mathscr{F}) is a τGξ\tau^{G\oplus\xi}-embedding of a manifold MnM^{n} into Pn+k+m×rP^{n+k+m}\times\mathbb{R}^{r}, then by the above lemma we obtain a diffeotopy of P×rP\times\mathbb{R}^{r} that turns 𝒱\mathscr{V} and \mathscr{F} vertical. A diffeotopy of P×rP\times\mathbb{R}^{r} also yields a cobordism of τGξ\tau^{G\oplus\xi}-embeddings, hence we can assume that 𝒱\mathscr{V} and \mathscr{F} were initially vertical. Now we can define β\beta to assign to the cobordism class of ee the cobordism class of the τGξ\tau^{G\oplus\xi}-map f~:ξMP\tilde{f}\colon\xi_{M}\to P for which f~|M=prPe\tilde{f}|_{M}=\mathop{\rm pr}_{P}\circ e and f~|ξM|p=prPexpp\tilde{f}|_{\xi_{M}|_{p}}=\mathop{\rm pr}_{P}\circ\exp_{p} for all pe(M)p\in e(M) (where prP\mathop{\rm pr}_{P} denotes the projection to PP and expp\exp_{p} now denotes the exponential of the leaf of \mathscr{F} at pp). In order to prove that β\beta is well-defined, we have to show that the cobordism class of f~\tilde{f} does not depend on the choice of the representative of the cobordism class [e][e].

Claim.  If e0=(e0,𝒱0,ξM0,0)e_{0}=(e_{0},\mathscr{V}_{0},\xi_{M_{0}},\mathscr{F}_{0}) and e1=(e1,𝒱1,ξM1,1)e_{1}=(e_{1},\mathscr{V}_{1},\xi_{M_{1}},\mathscr{F}_{1}) are cobordant τGξ\tau^{G\oplus\xi}-embeddings of the manifolds M0M_{0} and M1M_{1} respectively into P×rP\times\mathbb{R}^{r} and the above method assigns to them the τGξ\tau^{G\oplus\xi}-maps f~0:ξM0P\tilde{f}_{0}\colon\xi_{M_{0}}\to P and f~1:ξM1P\tilde{f}_{1}\colon\xi_{M_{1}}\to P respectively, then f0f1f_{0}\sim f_{1}.

Proof.  We applied a diffeotopy φti(t[0,1])\varphi_{t}^{i}\leavevmode\nobreak\ (t\in[0,1]) of P×r×{i}P\times\mathbb{R}^{r}\times\{i\} to turn the vector fields 𝒱i\mathscr{V}_{i} and foliation i\mathscr{F}_{i} vertical (for i=0,1i=0,1), this way we obtained the τGξ\tau^{G\oplus\xi}-map f~i:ξMiP×{i}\tilde{f}_{i}\colon\xi_{M_{i}}\to P\times\{i\}. If e0e_{0} and e1e_{1} are connected by a cobordism E=(E,𝒰,ξW,𝒢)E=(E,\mathscr{U},\xi_{W},\mathscr{G}), which is a τGξ\tau^{G\oplus\xi}-embedding of a manifold Wn+1W^{n+1} into P×r×[0,1]P\times\mathbb{R}^{r}\times[0,1], then we can apply (the relative version of) the above lemma to obtain a diffeotopy Φt(t[0,1])\Phi_{t}\leavevmode\nobreak\ (t\in[0,1]) of P×r×[0,1]P\times\mathbb{R}^{r}\times[0,1] that extends the given diffeotopies φt0\varphi_{t}^{0} and φt1\varphi_{t}^{1} on the boundary and turns the vector fields 𝒰\mathscr{U} and the foliation 𝒢\mathscr{G} vertical. Now composing EE with the final diffeomorphism Φ1\Phi_{1} and the projection to P×[0,1]P\times[0,1] as above, we obtain a τGξ\tau^{G\oplus\xi}-cobordism F~:ξWP×[0,1]\tilde{F}\colon\xi_{W}\to P\times[0,1] between f0f_{0} and f1f_{1} for which F~|W=prP×[0,1]Φ1E\tilde{F}|_{W}=\mathop{\rm pr}_{P\times[0,1]}\circ\Phi_{1}\circ E and F~|ξW|p=prP×[0,1]Φ1expp\tilde{F}|_{\xi_{W}|_{p}}=\mathop{\rm pr}_{P\times[0,1]}\circ\Phi_{1}\circ\exp_{p} for all pE(W)p\in E(W).  \diamond

The constructions of α\alpha and β\beta imply also that they are homomorphisms and by design β\beta is the inverse of α\alpha, hence they are both isomophisms between CobτGξ(P)\textstyle{\mathop{\rm Cob}}_{\tau}^{G\oplus\xi}(P) and EmbτGξ(n,P×r)\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r}).  \square

Now we can describe the classifying space XτGξX_{\tau}^{G\oplus\xi} for which we shall need the pullback of ξBG\xi\to BG by the fibration π:KτGBG\pi\colon K_{\tau}^{G}\to BG (see definition 3.1).

Theorem 6.9.

If τ\tau is a set of stable singularities, then we have XτGξΓT(ντGπξ)X_{\tau}^{G\oplus\xi}\cong\Gamma T(\nu_{\tau}^{G}\oplus\pi^{*}\xi).

Proof.  We need to prove that cobordisms of τGξ\tau^{G\oplus\xi}-maps to an arbitrary fixed manifold Pn+k+mP^{n+k+m} bijectively correspond to homotopy classes of maps 𝑃ΓT(ντGπξ)\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\to\Gamma T(\nu_{\tau}^{G}\oplus\pi^{*}\xi). Since we have CobτGξ(P)EmbτGξ(n,P×r)\textstyle{\mathop{\rm Cob}}_{\tau}^{G\oplus\xi}(P)\cong\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r}) and [𝑃,ΓT(ντGπξ)][Sr𝑃,SrT(ντGπξ)][\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\Gamma T(\nu_{\tau}^{G}\oplus\pi^{*}\xi)]\cong[S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi)] for sufficiently large numbers rr, it is enough to prove

EmbτGξ(n,P×r)[Sr𝑃,SrT(ντGπξ)]\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r})\cong[S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi)]

if rr is large enough. As in the proof of the previous theorem we will define two homomorphisms α:EmbτGξ(n,P×r)[Sr𝑃,SrT(ντGπξ)]\alpha\colon\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r})\to[S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi)] and β:[Sr𝑃,SrT(ντGπξ)]EmbτGξ(n,P×r)\beta\colon[S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi)]\to\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r}) which are each other’s inverses as we shall see.

I. Construction of α:EmbτGξ(n,P×r)[Sr𝑃,SrT(ντGπξ)]\alpha\colon\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r})\to[S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi)].

Let e=(e,𝒱,ξM,)e=(e,\mathscr{V},\xi_{M},\mathscr{F}) be a τGξ\tau^{G\oplus\xi}-embedding of a manifold MnM^{n} into P×rP\times\mathbb{R}^{r}. The normal bundle of ee is of the form νe=νξM\nu_{e}=\nu^{\prime}\oplus\xi_{M} and by definition it is induced from γk+rGξ|BG(k+r)\gamma_{k+r}^{G}\oplus\xi|_{BG(k+r)} by a map μe:MBG(k+r)\mu_{e}\colon M\to BG(k+r). The embedding ee can be written as the composition

M𝑖DξM𝑗DνeP×rM\mathop{\xrightarrow{\leavevmode\nobreak\ i\leavevmode\nobreak\ }}D\xi_{M}\mathop{\xrightarrow{\leavevmode\nobreak\ j\leavevmode\nobreak\ }}D\nu_{e}\hookrightarrow P\times\mathbb{R}^{r}

where the disk bundles DξMD\xi_{M} and DνeD\nu_{e} are viewed as closed embedded tubular neighbourhoods with j1(Sνe)=SξMj^{-1}(S\nu_{e})=S\xi_{M} and the fibres of DξMD\xi_{M} are identified with the intersections of the leaves of \mathscr{F} with the tubular neighbourhood DξMD\xi_{M}.

Now the vector fields 𝒱\mathscr{V} can be extended to DξMD\xi_{M} by translation along the fibres and if LL is the leaf of \mathscr{F} at a point pe(M)p\in e(M), then we can use its exponential to define a foliation of dimension rr on LL which is tangent to the extended 𝒱\mathscr{V} along the disk DpξMD_{p}\xi_{M}. Applying this for all points pe(M)p\in e(M) we obtain the structure of a τG\tau^{G}-embedding on jj. Hence there are inducing maps κj:DξMKτG\kappa_{j}\colon D\xi_{M}\to K_{\tau}^{G} and κ~j:DνeSrTντG\tilde{\kappa}_{j}\colon D\nu_{e}\to S^{r}T\nu_{\tau}^{G} that pull back jj from the embedding KτGSrTντGK_{\tau}^{G}\hookrightarrow S^{r}T\nu_{\tau}^{G} (again we are considering such a large number rr and such a finite dimensional approximation of the Kazarian space that the space SrTντGS^{r}T\nu_{\tau}^{G} exists).

The normal bundle of e(M)e(M) in the disk bundle DξMD\xi_{M} is induced by the μe\mu_{e} above and this all fits into a diagram

Dνe\textstyle{D\nu_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ~j\scriptstyle{\tilde{\kappa}_{j}}SrTντG\textstyle{S^{r}T\nu_{\tau}^{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tγk+rG\textstyle{T\gamma_{k+r}^{G}}Tξ|BG(k+r)\textstyle{T\xi|_{BG(k+r)}}DξM\textstyle{D\xi_{M}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}κj\scriptstyle{\kappa_{j}}KτG\textstyle{K_{\tau}^{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}BG(k+r)\textstyle{BG(k+r)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}P×r\textstyle{P\times\mathbb{R}^{r}}BG(k+r)\textstyle{BG(k+r)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μe\scriptstyle{\mu_{e}}i\scriptstyle{i}e\scriptstyle{e}

Since μe\mu_{e} induces the normal bundle of ee, we have in the above diagram μe=πκji\mu_{e}=\pi\circ\kappa_{j}\circ i, that is, the inducing map μe\mu_{e} factors through the Kazarian space KτGK_{\tau}^{G}. But then we have a diagram

μeξ=ξM\textstyle{\mu_{e}^{*}\xi=\xi_{M}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πξ\textstyle{\pi^{*}\xi\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ξ|BG(k+r)\textstyle{\xi|_{BG(k+r)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μeγk+rG=ν\textstyle{\mu_{e}^{*}\gamma_{k+r}^{G}=\nu^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πγk+rG=ντGεr\textstyle{\pi^{*}\gamma_{k+r}^{G}=\nu_{\tau}^{G}\oplus\varepsilon^{r}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γk+rG\textstyle{\gamma_{k+r}^{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κe\scriptstyle{\kappa_{e}}μe\scriptstyle{\mu_{e}}KτG\textstyle{K_{\tau}^{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}BG(k+r)\textstyle{BG(k+r)}

where all squares are pullback squares.

Considering now the sum γk+rξ|BG(k+r)\gamma_{k+r}\oplus\xi|_{BG(k+r)}, we get νe=κeπ(γk+rξ|BG(k+r))=κe(ντGεrπξ)\nu_{e}=\kappa_{e}^{*}\pi^{*}(\gamma_{k+r}\oplus\xi|_{BG(k+r)})=\kappa_{e}^{*}(\nu_{\tau}^{G}\oplus\varepsilon^{r}\oplus\pi^{*}\xi), hence the embedding of MM into P×rP\times\mathbb{R}^{r} is induced by a diagram

Sr𝑃\textstyle{S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ~e\scriptstyle{\tilde{\kappa}_{e}}SrT(ντGπξ)\textstyle{S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi)}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e\scriptstyle{e}κe\scriptstyle{\kappa_{e}}KτG\textstyle{K_{\tau}^{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

This way we can assign to the τGξ\tau^{G\oplus\xi}-embedding ee the map κ~e\tilde{\kappa}_{e} and since we can apply this construction to cobordisms as well, the assignment α[e]:=[κ~e]\alpha[e]:=[\tilde{\kappa}_{e}] is well-defined.

II. Construction of β:[Sr𝑃,SrT(ντGπξ)]EmbτGξ(n,P×r)\beta\colon[S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi)]\to\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r}).

Suppose we have a map κ~:Sr𝑃SrT(ντGπξ)\tilde{\kappa}\colon S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\to S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi) and put Mn:=κ~1(KτG)M^{n}:=\tilde{\kappa}^{-1}(K_{\tau}^{G}) and κ:=κ~|M\kappa:=\tilde{\kappa}|_{M}. Now the normal bundle of MM in P×rP\times\mathbb{R}^{r} is of the form νξM\nu^{\prime}\oplus\xi_{M} where ν\nu^{\prime} and ξM\xi_{M} are induced from γk+rG\gamma_{k+r}^{G} and ξ|BG(k+r)\xi|_{BG(k+r)} respectively by the map πκ:MBG(k+r)\pi\circ\kappa\colon M\to BG(k+r).

If we take the preimage of the disk bundle Dξ|BG(k+r)D(γk+rGξ|BG(k+r))D\xi|_{BG(k+r)}\subset D(\gamma_{k+r}^{G}\oplus\xi|_{BG(k+r)}) in P×rP\times\mathbb{R}^{r} (of a sufficiently small radius), then we get an embedding j:DξMDνMj\colon D\xi_{M}\hookrightarrow D\nu_{M} where νM\nu_{M} is the normal bundle of MM and the disk bundles are again considered as closed tubular neighbourhoods. Now we can project DνMD\nu_{M} to the orthogonal complement of DξMD\xi_{M} in it along the fibres of DξMD\xi_{M} and if we denote this projection by ρ\rho, then the composition κ~ρ\tilde{\kappa}\circ\rho is a map of DνMD\nu_{M} to SrTντGS^{r}T\nu_{\tau}^{G} such that the preimage of KτGK_{\tau}^{G} is DξMD\xi_{M}, hence it gives the structure of a τG\tau^{G}-embedding to jj.

We obtain vector fields 𝒱=(v1,,vr)\mathscr{V}=(v_{1},\ldots,v_{r}) along MM by restricting the vector fields along DξMD\xi_{M}. We also have a foliation 𝒢\mathscr{G} of dimension rr on a neighbourhood of MM that is tangent to 𝒱\mathscr{V} along MM and the projection ρ\rho maps each leaf of 𝒢\mathscr{G} again into a leaf of 𝒢\mathscr{G} because of the definition of the inducing map of the τG\tau^{G}-embedding jj. So we can define a foliation \mathscr{F} of dimension r+mr+m by defining the leaf of \mathscr{F} at a point pp to be the preimage of the leaf of 𝒢\mathscr{G} at pp under ρ\rho.

This yields the structure of a τGξ\tau^{G\oplus\xi}-embedding e=(e,𝒱,ξM,)e=(e,\mathscr{V},\xi_{M},\mathscr{F}) on the embedding of MM into P×rP\times\mathbb{R}^{r}. The same method assigns to a homotopy of κ~:Sr𝑃SrT(ντGπξ)\tilde{\kappa}\colon S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}}\to S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi) a cobordism of ee, hence we can define β[κ~]:=[e]\beta[\tilde{\kappa}]:=[e].

The above constructions imply that α\alpha and β\beta are homomorphisms and also that they are inverses of each other, hence we have proved EmbτGξ(n,P×r)[Sr𝑃,SrT(ντGπξ)]\textstyle{\mathop{\rm Emb}}_{\tau}^{G\oplus\xi}(n,P\times\mathbb{R}^{r})\cong[S^{r}\mathop{\mathop{P}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},S^{r}T(\nu_{\tau}^{G}\oplus\pi^{*}\xi)] and this is what we wanted.  \square

7 Another long exact sequence

We shall apply the results of the section above in the case G=OG=\mathrm{O} and ξ=2detγ=detγdetγ\xi=2\det\gamma=\det\gamma\oplus\det\gamma (recall that detγ\det\gamma was the line bundle over BOB\mathrm{O} which induces detγnO\det\gamma_{n}^{\mathrm{O}} over BO(n)B\mathrm{O}(n) for all nn). The pullback of detγ\det\gamma over the Kazarian space KτOK_{\tau}^{\mathrm{O}} is then detντO\det\nu_{\tau}^{\mathrm{O}} and so the classifying space of τO2detγ\tau^{\mathrm{O}\oplus 2\det\gamma}-cobordisms is ΓT(ντO2detντO)\Gamma T(\nu_{\tau}^{\mathrm{O}}\oplus 2\det\nu_{\tau}^{\mathrm{O}}).

Remark 7.1.

Note that informally a τO2detγ\tau^{\mathrm{O}\oplus 2\det\gamma}-map can be thought of as a “τ\tau-map” f:MnPn+k+2f\colon M^{n}\to P^{n+k+2} with normal bundle of the form νf=ν2detν\nu_{f}=\nu^{\prime}\oplus 2\det\nu^{\prime}. Now this detν\det\nu^{\prime} coincides with detνf\det\nu_{f} hence ff 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 τ\tau of stable singularities and any manifold QqQ^{q} there is a long exact sequence

\displaystyle\ldots ψm+1Cobτint(Q×m)φmCobτO(Q×m)χmCobτO2detγ(Q×m)ψm\displaystyle\mathop{\xrightarrow{\leavevmode\nobreak\ \psi^{\prime}_{m+1}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \varphi^{\prime}_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}^{\mathrm{O}}_{\tau}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \chi^{\prime}_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(Q\times\mathbb{R}^{m})\mathop{\xrightarrow{\leavevmode\nobreak\ \psi^{\prime}_{m}\leavevmode\nobreak\ }}
ψmCobτint(Q×m1)\displaystyle\mathop{\xrightarrow{\leavevmode\nobreak\ \psi^{\prime}_{m}\leavevmode\nobreak\ }}\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m-1})\to\ldots

Proof.  Recall the pullback diagram

Bint\textstyle{B\mathrm{int}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S1\textstyle{S^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{i}BO\textstyle{B\mathrm{O}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w1\scriptstyle{w_{1}}P\textstyle{\mathbb{R}P^{\infty}}

from definition 3.8. The complement of S1S^{1} in P\mathbb{R}P^{\infty} deformation retracts to P2\mathbb{R}P^{\infty-2} which is also a deformation retract of P\mathbb{R}P^{\infty}, hence by pulling back its embedding by the fibration w1:BOPw_{1}\colon B\mathrm{O}\to\mathbb{R}P^{\infty} to the embedding of a space BB we get a homotopy equivalence as shown on the diagram

BO\textstyle{B\mathrm{O}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w1\scriptstyle{w_{1}}P\textstyle{\mathbb{R}P^{\infty}}B\textstyle{B\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}P2\textstyle{\mathbb{R}P^{\infty-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}

Next we pull back the bundle KτOBOK_{\tau}^{\mathrm{O}}\to B\mathrm{O} by the homotopy equivalence BBOB\to B\mathrm{O} to a bundle KBK\to B, thus we have KKτOK\cong K_{\tau}^{\mathrm{O}}. Then the pullback of the universal virtual normal bundle ντO\nu_{\tau}^{\mathrm{O}} over KK, which will be denoted ν\nu, is stably isomorphic to ντO\nu_{\tau}^{\mathrm{O}}.

Take the virtual bundle ντint\nu_{\tau}^{\mathrm{int}} over KτintKτOK_{\tau}^{\mathrm{int}}\subset K_{\tau}^{\mathrm{O}} (which is the restriction of ντO\nu_{\tau}^{\mathrm{O}}) and the cofibration

SnTντint(n)SnTντO(n)SnTντO(n)/SnTντint(n)S^{n}T\nu_{\tau}^{\mathrm{int}}(n)\hookrightarrow S^{n}T\nu_{\tau}^{\mathrm{O}}(n)\to S^{n}T\nu_{\tau}^{\mathrm{O}}(n)/S^{n}T\nu_{\tau}^{\mathrm{int}}(n)

for any nn. The normal bundle of KKτOK\subset K_{\tau}^{\mathrm{O}} is induced from the normal bundle of P2P\mathbb{R}P^{\infty-2}\subset\mathbb{R}P^{\infty} by the composition of w1:BP2w_{1}\colon B\to\mathbb{R}P^{\infty-2} with the fibration KBK\to B. The normal bundle of P2\mathbb{R}P^{\infty-2} is 2γ1O2\gamma_{1}^{\mathrm{O}}, this induces 2detγ2\det\gamma over BB and this finally induces 2detν2\det\nu over the Kazarian space KK, hence the cofibration above has the form

SnTντint(n)SnTντO(n)SnT(ντO(n)2detντO(n)).S^{n}T\nu_{\tau}^{\mathrm{int}}(n)\hookrightarrow S^{n}T\nu_{\tau}^{\mathrm{O}}(n)\to S^{n}T(\nu_{\tau}^{\mathrm{O}}(n)\oplus 2\det\nu_{\tau}^{\mathrm{O}}(n)).

Now applying the functor Ωn+mΓ\Omega^{n+m}\Gamma to the Puppe sequence of this cofibration we get a sequence of maps

ΩmΓTντint(n)ΩmΓTντO(n)ΩmΓT(ντO(n)2detντO(n))Ωm1ΓTντint(n)\ldots\to\Omega^{m}\Gamma T\nu_{\tau}^{\mathrm{int}}(n)\to\Omega^{m}\Gamma T\nu_{\tau}^{\mathrm{O}}(n)\to\Omega^{m}\Gamma T(\nu_{\tau}^{\mathrm{O}}(n)\oplus 2\det\nu_{\tau}^{\mathrm{O}}(n))\to\Omega^{m-1}\Gamma T\nu_{\tau}^{\mathrm{int}}(n)\to\ldots

This sequence is infinite to the right by construction, but it is also infinite to the left since the number nn could be arbitrary and we get the same maps by applying Ωn+mΓ\Omega^{n+m}\Gamma to the nn’th suspensions as by applying Ωn+m+1Γ\Omega^{n+m+1}\Gamma to the (n+1)(n+1)’st suspensions. We can now converge with nn to infinity since the maps in this sequence commute with the natural maps induced by the inclusions Kτσ(n)Kτσ(n+1)K_{\tau}^{\sigma}(n)\subset K_{\tau}^{\sigma}(n+1) (for σ=int,O\sigma=\mathrm{int},\mathrm{O}) and so we get a sequence of maps

ΩmΓTντintΩmΓTντOΩmΓT(ντO2detντO)Ωm1ΓTντint\ldots\to\Omega^{m}\Gamma T\nu_{\tau}^{\mathrm{int}}\to\Omega^{m}\Gamma T\nu_{\tau}^{\mathrm{O}}\to\Omega^{m}\Gamma T(\nu_{\tau}^{\mathrm{O}}\oplus 2\det\nu_{\tau}^{\mathrm{O}})\to\Omega^{m-1}\Gamma T\nu_{\tau}^{\mathrm{int}}\to\ldots

If we then fix a manifold QQ and apply the functor [𝑄,][\mathop{\mathop{Q}\limits^{\vbox to-0.3014pt{\kern-1.20552pt\hbox{${}^{\bullet}$}\vss}}},\cdot] to this sequence, then we obtain the long exact sequence of cobordism groups as claimed.  \square

Remark 7.3.

Similarly to the oriented case, if we had Q=qQ=\mathbb{R}^{q}, then the same long exact sequence could be obtained by turning the cofibration

TντintTντOT(ντO2detντO)T\nu_{\tau}^{\mathrm{int}}\hookrightarrow T\nu_{\tau}^{\mathrm{O}}\to T(\nu_{\tau}^{\mathrm{O}}\oplus 2\det\nu_{\tau}^{\mathrm{O}})

(cf. remark 4.2) into the fibration

ΓTντOΓTντintΓT(ντO2detντO)\Gamma T\nu_{\tau}^{\mathrm{O}}\mathop{\xrightarrow{\leavevmode\nobreak\ \Gamma T\nu_{\tau}^{\mathrm{int}}\leavevmode\nobreak\ }}\Gamma T(\nu_{\tau}^{\mathrm{O}}\oplus 2\det\nu_{\tau}^{\mathrm{O}})

by the functor Γ\Gamma and taking its homotopy long exact sequence.

Observe that the restriction of 2γ1O2\gamma_{1}^{\mathrm{O}} over the 11-cell S1PS^{1}\subset\mathbb{R}P^{\infty} is trivial, hence the pullback of 2detγBO2\det\gamma\to B\mathrm{O} over BintB\mathrm{int} will also be trivial which then induces the trivial bundle over the Kazarian space KτintK_{\tau}^{\mathrm{int}} as well. Thus analogously to the proof of theorem 7.2 we get a cofibration

S2TντintT(ντO2detντO)T(ντO4detντO)S^{2}T\nu_{\tau}^{\mathrm{int}}\hookrightarrow T(\nu_{\tau}^{\mathrm{O}}\oplus 2\det\nu_{\tau}^{\mathrm{O}})\to T(\nu_{\tau}^{\mathrm{O}}\oplus 4\det\nu_{\tau}^{\mathrm{O}})

and iterating this process yields:

Corollary 7.4.

For any set τ\tau of stable singularities and any manifold QqQ^{q} and integer r0r\geq 0 there is a long exact sequence

\displaystyle\ldots Cobτint(Q×m2r)CobτO2rdetγ(Q×m)CobτO2(r+1)detγ(Q×m)\displaystyle\to\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m-2r})\to\textstyle{\mathop{\rm Cob}}^{\mathrm{O}\oplus 2r\det\gamma}_{\tau}(Q\times\mathbb{R}^{m})\to\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2(r+1)\det\gamma}(Q\times\mathbb{R}^{m})\to
Cobτint(Q×m2r1)\displaystyle\to\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m-2r-1})\to\ldots
Remark 7.5.

It would be tempting to try finding homomorphisms which complete the commutative diagram

\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cobτint(Q×m)\textstyle{\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}CobτO2rdetγ(Q×m+2r)\textstyle{\textstyle{\mathop{\rm Cob}}^{\mathrm{O}\oplus 2r\det\gamma}_{\tau}(Q\times\mathbb{R}^{m+2r})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}CobτO2(r+1)detγ(Q×m+2r)\textstyle{\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2(r+1)\det\gamma}(Q\times\mathbb{R}^{m+2r})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cobτint(Q×m)\textstyle{\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}CobτO2sdetγ(Q×m+2s)\textstyle{\textstyle{\mathop{\rm Cob}}^{\mathrm{O}\oplus 2s\det\gamma}_{\tau}(Q\times\mathbb{R}^{m+2s})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}CobτO2(s+1)detγ(Q×m+2s)\textstyle{\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2(s+1)\det\gamma}(Q\times\mathbb{R}^{m+2s})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots}

with the dashed arrows for different numbers rr and ss, however, it seems that such homomorphisms do not exist in general.

Now it remains from the proof of theorem 2 to describe the homomorphisms φm\varphi^{\prime}_{m} and χm\chi^{\prime}_{m} in the exact sequence in theorem 7.2.

8 Description of the homomorphisms φm\varphi^{\prime}_{m}, χm\chi^{\prime}_{m} and ψm\psi^{\prime}_{m}

Proposition 8.1.

φm:Cobτint(Q×m)CobτO(Q×m)\varphi^{\prime}_{m}\colon\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m})\to\textstyle{\mathop{\rm Cob}}^{\mathrm{O}}_{\tau}(Q\times\mathbb{R}^{m}) is the forgetful homomorphism that assigns to the cobordism class of a Wall map (f,w)(f,w) the unoriented cobordism class of ff.

Proof.  This follows immediately since the map between the classifying spaces is just the inclusion of ΓTντint\Gamma T\nu_{\tau}^{\mathrm{int}} into ΓTντO\Gamma T\nu_{\tau}^{\mathrm{O}} (see the proof of theorem 7.2).  \square

Proposition 8.2.

χm:CobτO(Q×m)CobτO2detγ(Q×m)\chi^{\prime}_{m}\colon\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}}(Q\times\mathbb{R}^{m})\to\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(Q\times\mathbb{R}^{m}) assigns to a cobordism class [f][f] the cobordism class of ff restricted to the Poincaré dual of w1(νf)2w_{1}(\nu_{f})^{2} which can be represented by a 22-codimensional submanifold uniquely up to cobordism, the restriction f|PD(w1(νf)2)f|_{\mathop{\rm PD}(w_{1}(\nu_{f})^{2})} has a normal O2detγ\mathrm{O}\oplus 2\det\gamma-structure and its cobordism class only depends on the class of ff, hence this assignment is well-defined.

Proof.  This is completely analogous to proposition 5.5. For an unoriented cobordism class [f]CobτO(Q×m)[f]\in\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}}(Q\times\mathbb{R}^{m}) represented by a τ\tau-map f:MQ×mf\colon M\to Q\times\mathbb{R}^{m} the Poincaré dual of

w1(νf):MPw_{1}(\nu_{f})\colon M\to\mathbb{R}P^{\infty}

is represented by the preimage w1(νf)1(P1)w_{1}(\nu_{f})^{-1}(\mathbb{R}P^{\infty-1}). Then choosing two distinct hyperplanes P11\mathbb{R}P^{\infty-1}_{1} and P21\mathbb{R}P^{\infty-1}_{2} in P\mathbb{R}P^{\infty} the Poincaré dual of w1(νf)2w_{1}(\nu_{f})^{2} is represented by

w1(νf)1(P11)w1(νf)1(P21)\displaystyle w_{1}(\nu_{f})^{-1}(\mathbb{R}P^{\infty-1}_{1})\cap w_{1}(\nu_{f})^{-1}(\mathbb{R}P^{\infty-1}_{2}) =w1(νf)1(P11P21)=\displaystyle=w_{1}(\nu_{f})^{-1}(\mathbb{R}P^{\infty-1}_{1}\cap\mathbb{R}P^{\infty-1}_{2})=
=w1(νf)1(P2).\displaystyle=w_{1}(\nu_{f})^{-1}(\mathbb{R}P^{\infty-2}).

Now the same argument as in the proof of proposition 5.5 yields that χm\chi^{\prime}_{m} assigns to the cobordism class of ff the class of f|w1(νf)1(P2)=f|PD(w1(νf)2)f|_{w_{1}(\nu_{f})^{-1}(\mathbb{R}P^{\infty-2})}=f|_{\mathop{\rm PD}(w_{1}(\nu_{f})^{2})}.  \square

This finishes the proof of theorem 2.

Remark 8.3.

Although there are no apparent interesting algebraic properties of the homomorphism ψm:CobτO2detγ(Q×m)Cobτint(Q×m1)\psi^{\prime}_{m}\colon\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(Q\times\mathbb{R}^{m})\to\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{m-1}), it has a nice geometric description. Let qq be the dimension of QQ and put n:=q+mkn:=q+m-k. Observe that we have

CobτO2detγ(Qq×m)\displaystyle\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(Q^{q}\times\mathbb{R}^{m}) EmbτO2detγ(n2,Q×m+r)\displaystyle\cong\textstyle{\mathop{\rm Emb}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(n-2,Q\times\mathbb{R}^{m+r})\cong
EmbντOεr2detντO(n2,Q×m+r)\displaystyle\cong\textstyle{\mathop{\rm Emb}}^{\nu_{\tau}^{\mathrm{O}}\oplus\varepsilon^{r}\oplus 2\det\nu_{\tau}^{\mathrm{O}}}(n-2,Q\times\mathbb{R}^{m+r})

and

Cobτint(Qq×m1)\displaystyle\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q^{q}\times\mathbb{R}^{m-1}) Embτint1(n1,Q×m+r)\displaystyle\cong\textstyle{\mathop{\rm Emb}}_{\tau}^{\mathrm{int}\oplus 1}(n-1,Q\times\mathbb{R}^{m+r})\cong
Embντintεr+1(n1,Q×m+r)\displaystyle\cong\textstyle{\mathop{\rm Emb}}^{\nu_{\tau}^{\mathrm{int}}\oplus\varepsilon^{r+1}}(n-1,Q\times\mathbb{R}^{m+r})

where rr is sufficiently large, moreover, ψm\psi^{\prime}_{m} is obtained as the boundary homomorphism in the Puppe sequence of classifying spaces. Hence we can apply lemma 5.1 by setting the KK, AA, ξ\xi, BB and ζ\zeta in the lemma to be KτOK_{\tau}^{\mathrm{O}}, KK, 2detν2\det\nu, KτintK_{\tau}^{\mathrm{int}} and ντOεr\nu_{\tau}^{\mathrm{O}}\oplus\varepsilon^{r} respectively. This yields that for any cobordism class [f~]CobτO2detγ(Qq×m)[\tilde{f}]\in\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(Q^{q}\times\mathbb{R}^{m}) its image ψm[f~]\psi^{\prime}_{m}[\tilde{f}] is represented by the mapping of SξMS\xi_{M} (that is, the circle bundle of the 2detγ2\det\gamma-part of the normal bundle of f~|M\tilde{f}|_{M}) by the restriction of a representative of f~\tilde{f}, together with its natural outward normal vector field.

Remark 8.4.

As in remark 5.6 consider again the case Q×m=n+kQ\times\mathbb{R}^{m}=\mathbb{R}^{n+k} where the codimension kk is large (compared to nn). Then for any (non-empty) singularity set τ\tau we have Cobτint(n+k)=𝔚n\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(\mathbb{R}^{n+k})=\mathfrak{W}_{n} and CobτO(n+k)=𝔑n\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}}(\mathbb{R}^{n+k})=\mathfrak{N}_{n}, moreover, we also have CobτO2detγ(n+k)=:CobτO2detγ(n+k)=𝔑n2\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(\mathbb{R}^{n+k})=:\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(\mathbb{R}^{n+k})=\mathfrak{N}_{n-2} since for any embedding i:Mn2n+ki\colon M^{n-2}\hookrightarrow\mathbb{R}^{n+k} with normal bundle νi=ν2detν\nu_{i}=\nu^{\prime}\oplus 2\det\nu^{\prime} this detν\det\nu^{\prime} coincides with detνi\det\nu_{i} which is isomorphic to detTM\det TM, so this normal structure only depends on MM and if kk is large enough, 2detTM2\det TM can be embedded into n+k\mathbb{R}^{n+k} uniquely up to isotopy.

Thus theorem 2 gives an exact sequence for which the portion where nn is sufficiently small (compared to kk) looks like

𝔚n𝔑n𝔑n2𝔚n1\ldots\to\mathfrak{W}_{n}\to\mathfrak{N}_{n}\to\mathfrak{N}_{n-2}\to\mathfrak{W}_{n-1}\to\ldots

and if we increase kk with a fixed nn, 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 kk was assumed to be large enough, the manifold constructed in [atiyah, theorem 4.4] can also be mapped to n+k\mathbb{R}^{n+k} uniquely up to isotopy which now gives a splitting 𝔑n2𝔑n\mathfrak{N}_{n-2}\to\mathfrak{N}_{n} 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 τ={Σ0}\tau=\{\Sigma^{0}\} (the case of immersions) and τ={Σ0,Σ1,0}\tau=\{\Sigma^{0},\Sigma^{1,0}\}, τ={Σ0,Σ1,0,Σ1,1,0}\tau=\{\Sigma^{0},\Sigma^{1,0},\Sigma^{1,1,0}\} and so on (the case of Morin maps). Our first goal will always be to describe how the endomorphism φm\varphi_{m} in theorem 2 acts rationally. By [hosszu, proposition 90] we have

CobτSO(Pn+k)i=1n+kHi(P;)Hik(KτSO;),\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(P^{n+k})\otimes\mathbb{Q}\cong\displaystyle\bigoplus_{i=1}^{n+k}H_{i}(P;\mathbb{Q})\otimes H^{i-k}(K_{\tau}^{\mathrm{SO}};\mathbb{Q}),

in particular CobτSO(n+k)Hn(KτSO;)\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(\mathbb{R}^{n+k})\otimes\mathbb{Q}\cong H^{n}(K_{\tau}^{\mathrm{SO}};\mathbb{Q}), hence in order to understand φm\varphi_{m} rationally we only have to know how the involution which induces ι\iota acts on the Kazarian space KτSOK_{\tau}^{\mathrm{SO}}.

After this rational description we shall compute in both cases some cobordism groups of Wall τ\tau-maps for the above singularity sets τ\tau 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 φm\varphi^{\prime}_{m} in theorem 2 is not always injective but otherwise we have the same property, namely for an unoriented τ\tau-cobordism class to be representable by an orientable τ\tau-map it should be the φm\varphi^{\prime}_{m}-image of a Wall τ\tau-cobordism class.

9 Immersions

In this section we investigate the case τ={Σ0}\tau=\{\Sigma^{0}\}, i.e. τ\tau-maps are the kk-codimensional immersions. We shall use the notation

Immσ(n,k):=Cob{Σ0}σ(n+k)andImmσ(n,Pn+k):=Cob{Σ0}σ(Pn+k)\textstyle{\mathop{\rm Imm}}^{\sigma}(n,k):=\textstyle{\mathop{\rm Cob}}_{\{\Sigma^{0}\}}^{\sigma}(\mathbb{R}^{n+k})\quad\text{and}\quad\textstyle{\mathop{\rm Imm}}^{\sigma}(n,P^{n+k}):=\textstyle{\mathop{\rm Cob}}_{\{\Sigma^{0}\}}^{\sigma}(P^{n+k})

for any stable normal normal structure σ\sigma and any manifold PP (except that for σ=Gξ\sigma=G\oplus\xi we decrease nn and increase kk by the rank of ξ\xi). Now the oriented Kazarian space is K{Σ0}SO=BSO(k)K_{\{\Sigma^{0}\}}^{\mathrm{SO}}=B\mathrm{SO}(k) and the universal normal bundle over it is γkSO\gamma_{k}^{\mathrm{SO}} and we immediately obtain:

Proposition 9.1.

The endomorphism φn+k\varphi_{n+k}\otimes\mathbb{Q} of ImmSO(n,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n,k)\otimes\mathbb{Q} is the following:

  1. (1)

    if k=2mk=2m is even and qq is the number of non-negative integers a0,a1,,ama_{0},a_{1},\ldots,a_{m} such that a0a_{0} is odd and n=a0k+𝑚i=1ai4in=a_{0}k+\mathop{\overset{m}{\sum}}\limits_{\vbox to-1.50694pt{\kern-4.82224pt\hbox{$\scriptstyle i=1$}\vss}}a_{i}4i, then φn+k\varphi_{n+k}\otimes\mathbb{Q} is trivial on qq generators and the multiplication by 22 on the rest of the generators (in an appropriate basis),

  2. (2)

    if kk is odd, then φn+k\varphi_{n+k}\otimes\mathbb{Q} is the multiplication by 22.

Proof.  We have

ImmSO(,k)\displaystyle\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(*,k)\otimes\mathbb{Q} H(BSO(k);)=\displaystyle\cong H^{*}(B\mathrm{SO}(k);\mathbb{Q})=
={[p1,,pm,e]/(e2pm),if k=2m is even[p1,,pm],if k=2m+1 is odd\displaystyle=\begin{cases}\mathbb{Q}[p_{1},\ldots,p_{m},e]/(e^{2}-p_{m}),&\text{if }k=2m\text{ is even}\\ \mathbb{Q}[p_{1},\ldots,p_{m}],&\text{if }k=2m+1\text{ is odd}\end{cases}

where the piH4i(BSO(k);)p_{i}\in H^{4i}(B\mathrm{SO}(k);\mathbb{Q}) are the Pontryagin classes and eHk(BSO;)e\in H^{k}(B\mathrm{SO};\mathbb{Q}) is the Euler class of γkSO\gamma_{k}^{\mathrm{SO}}. The identification of ImmSO(n,k)πn+k+1s(STγkSO)\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n,k)\otimes\mathbb{Q}\cong\pi^{s}_{n+k+1}(ST\gamma_{k}^{\mathrm{SO}}) with the degree-nn 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 γkSOε1BSO(k)\gamma_{k}^{\mathrm{SO}}\oplus\varepsilon^{1}\to B\mathrm{SO}(k).

By proposition 5.3 (see also corollary 5.2) the involution inducing ι\iota acts on this vector bundle so that it inverts the summand ε1\varepsilon^{1} and reverses orientation on the summand γkSO\gamma_{k}^{\mathrm{SO}}, thus the Pontryagin classes pip_{i} are unchanged by its action and the Euler class ee is mapped to e-e. Our claim easily follows from this.  \square

We can say more about these cobordism groups in those cases where the codimension kk is either small or relatively large.

Small codimensional immersions

First, if the codimension is 11, then the normal bundle of any immersion f:MnPn+1f\colon M^{n}\looparrowright P^{n+1} is induced by w1(νf):MnPw_{1}(\nu_{f})\colon M^{n}\to\mathbb{R}P^{\infty} from γ1O\gamma_{1}^{\mathrm{O}}. The classifying spaces of ImmSO\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}, Immint\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}, ImmO\textstyle{\mathop{\rm Imm}}^{\mathrm{O}} and ImmO2detγ\textstyle{\mathop{\rm Imm}}^{\mathrm{O}\oplus 2\det\gamma} are ΓTε1\Gamma T\varepsilon^{1}, ΓTγ1O|P1\Gamma T\gamma^{\mathrm{O}}_{1}|_{\mathbb{R}P^{1}}, ΓTγ1O\Gamma T\gamma^{\mathrm{O}}_{1} and ΓT(3γ1O)\Gamma T(3\gamma^{\mathrm{O}}_{1}) respectively and since we have T(kγ1O|Pm)=Pm+k/Pk1T(k\gamma_{1}^{\mathrm{O}}|_{\mathbb{R}P^{m}})=\mathbb{R}P^{m+k}/\mathbb{R}P^{k-1} (for all kk and mm) these classifying spaces are ΓS1\Gamma S^{1}, ΓP2\Gamma\mathbb{R}P^{2}, ΓP\Gamma\mathbb{R}P^{\infty} and Γ(P/P2)\Gamma(\mathbb{R}P^{\infty}/\mathbb{R}P^{2}) respectively.

Remark 9.2.

The group ImmO2detγ(n2,Pn+1)\textstyle{\mathop{\rm Imm}}^{\mathrm{O}\oplus 2\det\gamma}(n-2,P^{n+1}) is the cobordism group of those immersions f:Mn2Pn+1f\colon M^{n-2}\looparrowright P^{n+1} whose normal bundle splits to the sum of three identical line bundles, i.e. the group Immsfr(n2,P)\textstyle{\mathop{\rm Imm}}^{\text{sfr}}(n-2,P) of 33-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

πs(n)πs(n)πn+1s(P2)πs(n1)\ldots\to\pi^{s}(n)\to\pi^{s}(n)\to\pi^{s}_{n+1}(\mathbb{R}P^{2})\to\pi^{s}(n-1)\to\ldots

and

πn+1s(P2)πn+1s(P)πn+1s(P/P2)πns(P2)\ldots\to\pi^{s}_{n+1}(\mathbb{R}P^{2})\to\pi^{s}_{n+1}(\mathbb{R}P^{\infty})\to\pi^{s}_{n+1}(\mathbb{R}P^{\infty}/\mathbb{R}P^{2})\to\pi^{s}_{n}(\mathbb{R}P^{2})\to\ldots

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.

The composition ψmψm+1\psi_{m}\circ\psi^{\prime}_{m+1} of the homomorphisms ψm+1\psi^{\prime}_{m+1} in theorem 2 and ψm\psi_{m} in theorem 2 is not always zero.

This is interesting since the classical analogue of ψm+1\psi^{\prime}_{m+1} 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 11-codimensional immersions of 44-manifolds. Then the combination of the above two exact sequences gives a diagram

π6s(P/P2)\textstyle{\pi^{s}_{6}(\mathbb{R}P^{\infty}/\mathbb{R}P^{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ6\scriptstyle{\psi^{\prime}_{6}}πs(4)\textstyle{\pi^{s}(4)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π5s(P2)\textstyle{\pi^{s}_{5}(\mathbb{R}P^{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ5\scriptstyle{\psi_{5}}πs(3)\textstyle{\pi^{s}(3)}π5s(P)\textstyle{\pi^{s}_{5}(\mathbb{R}P^{\infty})}

with exact row and column. Since πs(4)\pi^{s}(4) and π5s(P)\pi^{s}_{5}(\mathbb{R}P^{\infty}) are trivial (see [liu]) we get that ψ6\psi^{\prime}_{6} is epi and ψ5\psi_{5} is mono and since π5s(P2)\pi^{s}_{5}(\mathbb{R}P^{2}) is 2\mathbb{Z}_{2} (see [wu]) this means that ψ5ψ6\psi_{5}\circ\psi^{\prime}_{6} is non-zero.  \square

Now let us consider immersions of codimension 22. In this case the classifying space of ImmSO\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}} is ΓTγ2SO\Gamma T\gamma_{2}^{\mathrm{SO}} and since γ2SO\gamma_{2}^{\mathrm{SO}} coincides with the tautological complex line bundle over P\mathbb{C}P^{\infty} this space is ΓP\Gamma\mathbb{C}P^{\infty}. The first few stable homotopy groups of P\mathbb{C}P^{\infty} were computed by Liulevicius [liu] and Mosher [mosh] and are as follows:

mm 11 22 33 44 55 66 77 88 99 1010 1111 1212
πms(P)\pi^{s}_{m}(\mathbb{C}P^{\infty}) 0 \mathbb{Z} 0 \mathbb{Z} 2\mathbb{Z}_{2} \mathbb{Z} 2\mathbb{Z}_{2} 2\mathbb{Z}\oplus\mathbb{Z}_{2} 83\mathbb{Z}_{8}\oplus\mathbb{Z}_{3} \mathbb{Z} 4\mathbb{Z}_{4} 3\mathbb{Z}\oplus\mathbb{Z}_{3}

The action of the involution ι\iota in theorem 2 for ImmSO(n,2)πn+2s(P)\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n,2)\cong\pi^{s}_{n+2}(\mathbb{C}P^{\infty}) now immediately follows for n6n\leq 6 from proposition 9.1, hence we have:

Proposition 9.4.

The endomorphism φn+2\varphi_{n+2} of ImmSO(n,2)\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n,2) is

  1. (1)

    0 on the free part for all n2(4)n\equiv 2\leavevmode\nobreak\ (4) and 2id2\mathop{\rm id} on the free part for all n0(4)n\equiv 0\leavevmode\nobreak\ (4),

  2. (2)

    0 on the torsion part for 0n60\leq n\leq 6.

Corollary 9.5.

The cobordism groups Immint(n,2)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(n,2) for n6n\leq 6 are

nn 0 11 22 33 44 55 66
Immint(n,2)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(n,2) 2\mathbb{Z}_{2} 0 \mathbb{Z} 2\mathbb{Z}\oplus\mathbb{Z}_{2} 2?2\mathbb{Z}_{2}\leavevmode\nobreak\ ?\leavevmode\nobreak\ \mathbb{Z}_{2} 2\mathbb{Z}_{2} (2?2)\mathbb{Z}\oplus(\mathbb{Z}_{2}\leavevmode\nobreak\ ?\leavevmode\nobreak\ \mathbb{Z}_{2})

where G?HG\leavevmode\nobreak\ ?\leavevmode\nobreak\ H denotes the existence of a short exact sequence 0GG?HH00\to G\to G\leavevmode\nobreak\ ?\leavevmode\nobreak\ H\to H\to 0.

Proof.  Use theorem 2 to obtain the exact sequence

0cokerφn+2Immint(n,2)kerφn+100\to\mathop{\rm coker}\varphi_{n+2}\to\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(n,2)\to\ker\varphi_{n+1}\to 0

and apply the proposition above.  \square

Large codimensional immersions

Let us now turn to cobordisms of kk-codimensional immersions such that the dimension of the source manifolds is not much greater than kk. The reason for this is that for nn close to kk the group Immσ(n,k)\textstyle{\mathop{\rm Imm}}^{\sigma}(n,k) is close to the abstract cobordism group of nn-manifolds with normal σ\sigma-structures; see the works of Koschorke [kos], Olk [olk], Pastor [pas] and Li [li] where these groups were computed for knk+2k\leq n\leq k+2 and σ=O,SO\sigma=\mathrm{O},\mathrm{SO}.

Remark 9.6.

For n<kn<k we have ImmSO(n,k)=Ωn\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n,k)=\Omega_{n} and ImmO(n,k)=𝔑n\textstyle{\mathop{\rm Imm}}^{\mathrm{O}}(n,k)=\mathfrak{N}_{n} since a cobordism between two nn-manifolds is at most a kk-manifold which, when generically mapped with codimension kk, is immersed.

We shall use the natural forgetful homomorphisms

αimmSO:ImmSO(n,k)Ωn,αimmint:Immint(n,k)𝔚n,\displaystyle\alpha^{\mathrm{SO}}_{\mathrm{imm}}\colon\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n,k)\to\Omega_{n},\quad\alpha^{\mathrm{int}}_{\mathrm{imm}}\colon\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(n,k)\to\mathfrak{W}_{n},
αimmO:ImmO(n,k)𝔑nandαimmO2detγ:ImmO2detγ(n2,k+2)𝔑n2\displaystyle\alpha^{\mathrm{O}}_{\mathrm{imm}}\colon\textstyle{\mathop{\rm Imm}}^{\mathrm{O}}(n,k)\to\mathfrak{N}_{n}\quad\text{and}\quad\alpha^{\mathrm{O}\oplus 2\det\gamma}_{\mathrm{imm}}\colon\textstyle{\mathop{\rm Imm}}^{\mathrm{O}\oplus 2\det\gamma}(n-2,k+2)\to\mathfrak{N}_{n-2}

that assign to the cobordism class of an immersion f:Mnn+kf\colon M^{n}\looparrowright\mathbb{R}^{n+k} (or a germ f~:ξMn+k\tilde{f}\colon\xi_{M}\to\mathbb{R}^{n+k}) with the appropriate normal structure the abstract cobordism class of MM. 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

\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ImmSO(n,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmSO\scriptstyle{\alpha^{\mathrm{SO}}_{\mathrm{imm}}}ImmSO(n,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmSO\scriptstyle{\alpha^{\mathrm{SO}}_{\mathrm{imm}}}Immint(n,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{imm}}}ImmSO(n1,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n-1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmSO\scriptstyle{\alpha^{\mathrm{SO}}_{\mathrm{imm}}}\textstyle{\ldots}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωn\textstyle{\Omega_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2id\scriptstyle{2\mathop{\rm id}}Ωn\textstyle{\Omega_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚n\textstyle{\mathfrak{W}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωn1\textstyle{\Omega_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2id\scriptstyle{2\mathop{\rm id}}\textstyle{\ldots}

and

\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Immint(n,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{imm}}}ImmO(n,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{O}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmO\scriptstyle{\alpha^{\mathrm{O}}_{\mathrm{imm}}}ImmO2detγ(n2,k+2)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{O}\oplus 2\det\gamma}(n-2,k+2)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmO2detγ\scriptstyle{\alpha^{\mathrm{O}\oplus 2\det\gamma}_{\mathrm{imm}}}Immint(n1,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(n-1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{imm}}}\textstyle{\ldots}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\scriptstyle{0}𝔚n\textstyle{\mathfrak{W}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔑n\textstyle{\mathfrak{N}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔑n2\textstyle{\mathfrak{N}_{n-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\scriptstyle{0}𝔚n1\textstyle{\mathfrak{W}_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots}

where the rows are exact.

Lemma 9.7.

The map αimmO2detγ\alpha^{\mathrm{O}\oplus 2\det\gamma}_{\mathrm{imm}} is an isomorphism ImmO2detγ(n2,k+2)𝔑n2\textstyle{\mathop{\rm Imm}}^{\mathrm{O}\oplus 2\det\gamma}(n-2,k+2)\cong\mathfrak{N}_{n-2} for all nk+1n\leq k+1.

Proof.  We define the inverse of αimmO2detγ\alpha^{\mathrm{O}\oplus 2\det\gamma}_{\mathrm{imm}} in the following way: for a cobordism class [M]𝔑n2[M]\in\mathfrak{N}_{n-2} represented by Mn2M^{n-2} take the bundle ξM:=2detTM\xi_{M}:=2\det TM and map it generically to n+k\mathbb{R}^{n+k}; let the germ of this map along MM be f~\tilde{f} and assign the cobordism class [f~][\tilde{f}] to [M][M]. Such a correspondence would be the inverse of αimmO2detγ\alpha^{\mathrm{O}\oplus 2\det\gamma}_{\mathrm{imm}} if it was well-defined since detTM\det TM is also the determinant bundle of the stabilisation of ν(f~|M)\nu_{(\tilde{f}|_{M})}. So we only have to prove that f~\tilde{f} is an immersion germ and its cobordism class only depends on that of MM.

Let Nn2N^{n-2} be another representative of [M][M] and let Wn1W^{n-1} be a compact manifold with boundary such that W=MN\partial W=M\sqcup N. Then the restriction of ξW=2detTW\xi_{W}=2\det TW to the boundary gives ξM=2detTM\xi_{M}=2\det TM and ξN=2detTN\xi_{N}=2\det TN and if f~\tilde{f} and g~\tilde{g} are their generic map germs to n+k\mathbb{R}^{n+k} as defined above, then we can extend their representatives f:ξMn+k×{0}f\colon\xi_{M}\to\mathbb{R}^{n+k}\times\{0\} and g:ξNn+k×{1}g\colon\xi_{N}\to\mathbb{R}^{n+k}\times\{1\} by a generic map

F:ξWn+k×[0,1].F\colon\xi_{W}\to\mathbb{R}^{n+k}\times[0,1].

By the dimension condition such a map is stable. Its singular locus Σ(F):={pWrkdFp<n1}\Sigma(F):=\{p\in W\mid\mathop{\rm rk}dF_{p}<n-1\} is a submanifold of ξW\xi_{W} and a straightforward computation with jet bundles shows that we have codimΣ(F)=k+1\mathop{\rm codim}\Sigma(F)=k+1. Now since the dimension of WW is at most kk we generically have WΣ(F)=W\cap\Sigma(F)=\varnothing, that is, the map FF in a small neighbourhood of the zero-section is an immersion. This means that the germ F~\tilde{F} of FF along WW satisfies both conditions in definition 6.1, hence both f~\tilde{f} and g~\tilde{g} are O2detγ\mathrm{O}\oplus 2\det\gamma-immersions and F~\tilde{F} is an immersed O2detγ\mathrm{O}\oplus 2\det\gamma-cobordism between them. This finishes the proof.  \square

Proposition 9.8.

For all k1k\geq 1 we have

  1. (1)

    Immint(k,k)𝔚k\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k,k)\cong\mathfrak{W}_{k}\oplus\mathbb{Z} if kk is even,

  2. (2)

    Immint(k,k)𝔚k2\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k,k)\cong\mathfrak{W}_{k}\oplus\mathbb{Z}_{2} if kk is odd.

Proof.  By Koschorke [kos, theorem 10.8] and Pastor [pas, theorem 3.1] we have

ImmO(k,k)=𝔑kGandImmSO(k,k)=ΩkG\textstyle{\mathop{\rm Imm}}^{\mathrm{O}}(k,k)=\mathfrak{N}_{k}\oplus G\quad\text{and}\quad\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(k,k)=\Omega_{k}\oplus G

where GG denotes \mathbb{Z} if kk is even and 2\mathbb{Z}_{2} if kk is odd. Now by proposition 9.1 and lemma 9.7 the two diagrams above lemma 9.7 take the form

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk2ΩkG\textstyle{\frac{\Omega_{k}}{2\Omega_{k}}\oplus G\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}Immint(k,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{imm}}}T2(Ωk1)\textstyle{T_{2}(\Omega_{k-1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk2Ωk\textstyle{\frac{\Omega_{k}}{2\Omega_{k}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚k\textstyle{\mathfrak{W}_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T2(Ωk1)\textstyle{T_{2}(\Omega_{k-1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

and

Immint(n,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{imm}}}𝔑kG\textstyle{\mathfrak{N}_{k}\oplus G\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}𝔑k2\textstyle{\mathfrak{N}_{k-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚k\textstyle{\mathfrak{W}_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔑k\textstyle{\mathfrak{N}_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔑k2\textstyle{\mathfrak{N}_{k-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

where T2(Ωk1)T_{2}(\Omega_{k-1}) is the 22-torsion subgroup of Ωk1\Omega_{k-1}. The 55-lemma (or the snake lemma) applied to the first of these diagrams implies that αimmint\alpha^{\mathrm{int}}_{\mathrm{imm}} is epi and its kernel is GG and the second diagram implies that this kernel is also a direct summand.  \square

Remark 9.9.

Actually the proposition above could also be proved without using theorems 2 and 2 just by noting that the kernels of αimmSO\alpha^{\mathrm{SO}}_{\mathrm{imm}} and αimmO\alpha^{\mathrm{O}}_{\mathrm{imm}} are generated by the same object: the double point set (up to cobordism) of an immersion of SkS^{k} with one double point.

Proposition 9.10.

For all k1k\geq 1 we have

  1. (1)

    Immint(k+1,k)𝔚k+14\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k)\cong\mathfrak{W}_{k+1}\oplus\mathbb{Z}_{4} if k1(4)k\equiv 1\leavevmode\nobreak\ (4),

  2. (2)

    Immint(k+1,k)𝔚k+12\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k)\cong\mathfrak{W}_{k+1}\oplus\mathbb{Z}\oplus\mathbb{Z}_{2} if k2(4)k\equiv 2\leavevmode\nobreak\ (4),

  3. (3)

    Immint(k+1,k)𝔚k+12\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k)\cong\mathfrak{W}_{k+1}\oplus\mathbb{Z}_{2} if k3(4)k\equiv 3\leavevmode\nobreak\ (4),

  4. (4)

    Immint(k+1,k)𝔚k+122\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k)\cong\mathfrak{W}_{k+1}\oplus\mathbb{Z}\oplus\mathbb{Z}_{2}\oplus\mathbb{Z}_{2} if k0(4)k\equiv 0\leavevmode\nobreak\ (4).

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 kerαimmintkerαimmO\ker\alpha^{\mathrm{int}}_{\mathrm{imm}}\twoheadrightarrow\ker\alpha^{\mathrm{O}}_{\mathrm{imm}}. Moreover, if k+1k+1 is not a power of 22 and k1k\neq 1, then by [kos, theorem 10.8] kerαimmO\ker\alpha^{\mathrm{O}}_{\mathrm{imm}} is a direct summand in ImmO(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{O}}(k+1,k) and so if the image of [f]kerαimmint[f]\in\ker\alpha^{\mathrm{int}}_{\mathrm{imm}} is non-zero in kerαimmO\ker\alpha^{\mathrm{O}}_{\mathrm{imm}}, then [f][f] is independent in Immint(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k) of the subgroup generated by the αimmint\alpha^{\mathrm{int}}_{\mathrm{imm}}-preimage of 𝔚k+1{0}\mathfrak{W}_{k+1}\setminus\{0\}.

In the following we shall always use the first diagram above lemma 9.7 which yields (again by the 55-lemma) that αimmint\alpha^{\mathrm{int}}_{\mathrm{imm}} is epi in all cases. We will also use that by proposition 9.1 the endomorphism φ2k\varphi_{2k} is 0 on the direct complement of Ωk\Omega_{k} in ImmSO(k,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(k,k) and we obtain the forms of ImmSO(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(k+1,k) and ImmO(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{O}}(k+1,k) from [li, theorem 6] and [kos, theorem 10.8] respectively.

Proof of (1).  If k1(4)k\equiv 1\leavevmode\nobreak\ (4), we have

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk+12\textstyle{\Omega_{k+1}\oplus\mathbb{Z}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}Immint(k+1,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{imm}}}Ωk2\textstyle{\Omega_{k}\oplus\mathbb{Z}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk+1\textstyle{\Omega_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚k+1\textstyle{\mathfrak{W}_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk\textstyle{\Omega_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

hence kerαimmint\ker\alpha^{\mathrm{int}}_{\mathrm{imm}} is either 4\mathbb{Z}_{4} or 22\mathbb{Z}_{2}\oplus\mathbb{Z}_{2}. But kerαimmint\ker\alpha^{\mathrm{int}}_{\mathrm{imm}} maps onto kerαimmO4\ker\alpha^{\mathrm{O}}_{\mathrm{imm}}\cong\mathbb{Z}_{4}, so it can only be 4\mathbb{Z}_{4} which can only be a direct summand.

Proof of (2).  If k2(4)k\equiv 2\leavevmode\nobreak\ (4) and k+2k+2 is not a power of 22, then we have ImmSO(k+1,k)Ωk+14\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(k+1,k)\cong\Omega_{k+1}\oplus\mathbb{Z}_{4} and by [li, p. 472] the involution ι\iota in theorem 2 is the identity, hence the cokernel of φ2k+1\varphi_{2k+1} is Ωk+12\Omega_{k+1}\oplus\mathbb{Z}_{2}. If k+2k+2 is a power of 22, then we have ImmSO(k+1,k)Ωk+12\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(k+1,k)\cong\Omega_{k+1}\oplus\mathbb{Z}_{2} and so in both cases we get the diagram

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk+12\textstyle{\Omega_{k+1}\oplus\mathbb{Z}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}Immint(k+1,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{imm}}}Ωk\textstyle{\Omega_{k}\oplus\mathbb{Z}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk+1\textstyle{\Omega_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚k+1\textstyle{\mathfrak{W}_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk\textstyle{\Omega_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

which implies that we have kerαimmint2\ker\alpha^{\mathrm{int}}_{\mathrm{imm}}\cong\mathbb{Z}\oplus\mathbb{Z}_{2}. The \mathbb{Z} part cannot be anything else than a direct summand in Immint(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k) but the 2\mathbb{Z}_{2} part is also a direct summand since by the proof of [pas, theorem 3.2] it is generated by a cobordism class [f][f] which maps to the generator of the 2\mathbb{Z}_{2} part of ImmSO(k+1,k+1)Ωk+12\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(k+1,k+1)\cong\Omega_{k+1}\oplus\mathbb{Z}_{2} under the homomorphism induced by the inclusion 2k+12k+2\mathbb{R}^{2k+1}\subset\mathbb{R}^{2k+2} and this is non-zero in ImmO(k+1,k+1)\textstyle{\mathop{\rm Imm}}^{\mathrm{O}}(k+1,k+1) as well, hence the class of ff in ImmO(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{O}}(k+1,k) cannot vanish either.

Proof of (3).  If k3(4)k\equiv 3\leavevmode\nobreak\ (4), we have

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk+12Ωk+1\textstyle{\frac{\Omega_{k+1}}{2\Omega_{k+1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}Immint(k+1,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{imm}}}Ωk2\textstyle{\Omega_{k}\oplus\mathbb{Z}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk+12Ωk+1\textstyle{\frac{\Omega_{k+1}}{2\Omega_{k+1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚k+1\textstyle{\mathfrak{W}_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk\textstyle{\Omega_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

hence kerαimmint\ker\alpha^{\mathrm{int}}_{\mathrm{imm}} is 2\mathbb{Z}_{2} which can only be a direct summand (even if k+1k+1 is a power of 22).

Proof of (4).  If k0(4)k\equiv 0\leavevmode\nobreak\ (4), we have

Ωk+122\textstyle{\Omega_{k+1}\oplus\mathbb{Z}_{2}\oplus\mathbb{Z}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id+ι\scriptstyle{\mathop{\rm id}+\iota}Ωk+122\textstyle{\Omega_{k+1}\oplus\mathbb{Z}_{2}\oplus\mathbb{Z}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id00\scriptstyle{\mathop{\rm id}\oplus 0\oplus 0}Immint(k+1,k)\textstyle{\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αimmint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{imm}}}T2(Ωk)\textstyle{T_{2}(\Omega_{k})\oplus\mathbb{Z}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωk+1\textstyle{\Omega_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚k+1\textstyle{\mathfrak{W}_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T2(Ωk)\textstyle{T_{2}(\Omega_{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

and so kerαimmint\ker\alpha^{\mathrm{int}}_{\mathrm{imm}} is either 2\mathbb{Z}\oplus\mathbb{Z}_{2} or 22\mathbb{Z}\oplus\mathbb{Z}_{2}\oplus\mathbb{Z}_{2} depending on whether the ι\iota in theorem 2 swaps the two 2\mathbb{Z}_{2} summands in Ωk+122\Omega_{k+1}\oplus\mathbb{Z}_{2}\oplus\mathbb{Z}_{2} or not. Now again the proof of [pas, theorem 3.2] shows that one of these summands 2\mathbb{Z}_{2} is generated by a cobordism class [f][f] which maps to the generator of the 2\mathbb{Z}_{2} part of ImmSO(k+1,k+1)Ωk+12\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(k+1,k+1)\cong\Omega_{k+1}\oplus\mathbb{Z}_{2} under the homomorphism induced by the inclusion 2k+12k+2\mathbb{R}^{2k+1}\subset\mathbb{R}^{2k+2}. Since the reflection to a hyperplane commutes with this inclusion we get ι[f]=[f]\iota[f]=[f] which means that ι\iota does not swap the summands.

Hence we have kerαimmint22\ker\alpha^{\mathrm{int}}_{\mathrm{imm}}\cong\mathbb{Z}\oplus\mathbb{Z}_{2}\oplus\mathbb{Z}_{2} and we can also see that \mathbb{Z} is a direct summand in Immint(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k). Now if one of the 2\mathbb{Z}_{2}’s was not a direct summand in Immint(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k), that would mean that its generator coincided with 2[f]2[f] where [f][f] is in the preimage under αimmint\alpha^{\mathrm{int}}_{\mathrm{imm}} of a non-zero element in 𝔚k+1\mathfrak{W}_{k+1}, i.e. ff is an immersion Mk+12k+1M^{k+1}\looparrowright\mathbb{R}^{2k+1} where [M]𝔚k+1[M]\in\mathfrak{W}_{k+1} is not zero. This would imply that all elements in the preimage of [M][M] were of order 44 or \infty, hence to see that both 2\mathbb{Z}_{2}’s are direct summands it is sufficient to prove that the preimage under αimmint\alpha^{\mathrm{int}}_{\mathrm{imm}} of any element in 𝔚k+1\mathfrak{W}_{k+1} contains an element of order 22.

By [wallcob, section 4] we have that the 2\mathbb{Z}_{2}-algebra 𝔚\mathfrak{W}_{*} is generated by cobordism classes [P(2r1,2rs)][P(2^{r}-1,2^{r}s)], [Q(2r1,2rs)][Q(2^{r}-1,2^{r}s)] and [P2r][\mathbb{C}P^{2^{r}}] (for r,s1r,s\geq 1) of dimensions 2r(2s+1)12^{r}(2s+1)-1, 2r(2s+1)2^{r}(2s+1) and 2r+12^{r+1} respectively, moreover, P(2r1,2rs)P(2^{r}-1,2^{r}s) and P2r\mathbb{C}P^{2^{r}} are orientable while Q(2r1,2rs)Q(2^{r}-1,2^{r}s) is not cobordant to orientable manifolds. The monomials of rank k+1k+1 formed by these generators are a basis of the 2\mathbb{Z}_{2}-vector space 𝔚k+1\mathfrak{W}_{k+1}. We also have 𝔚k+1Ωk+1T2(Ωk)\mathfrak{W}_{k+1}\cong\Omega_{k+1}\oplus T_{2}(\Omega_{k}) and in the above basis the summands Ωk+1\Omega_{k+1} and T2(Ωk)T_{2}(\Omega_{k}) are generated respectively by the monomials containing an even and an odd number of [Q(2r1,2rs)][Q(2^{r}-1,2^{r}s)]’s since Ωk+1\Omega_{k+1} contains all classes representable by orientable manifolds.

Now Ωk+1𝔚k+1\Omega_{k+1}\subset\mathfrak{W}_{k+1} is independent of the 2\mathbb{Z}_{2}’s in Immint(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k) which means that all basis elements of 𝔚k+1\mathfrak{W}_{k+1} which are representable by orientable manifolds have elements of order 22 in their αimmint\alpha^{\mathrm{int}}_{\mathrm{imm}}-preimage. Thus it is enough to prove that this also holds for the unorientable basis elements, that is, any monomial of rank k+1k+1 formed by the manifolds P(2r1,2rs)P(2^{r}-1,2^{r}s), Q(2r1,2rs)Q(2^{r}-1,2^{r}s) and P2r\mathbb{C}P^{2^{r}} such that it contains an odd number of Q(2r1,2rs)Q(2^{r}-1,2^{r}s)’s can be immersed into 2k+1\mathbb{R}^{2k+1} representing a cobordism class of order 22.

Observe that Q(2r1,2rs)Q(2^{r}-1,2^{r}s) is even dimensional for all r,s1r,s\geq 1 and k+1k+1 is odd, hence such a monomial can be written in the form Mm×NnM^{m}\times N^{n} with m,n1m,n\geq 1 and m+n=k+1m+n=k+1. We can assume that mm is even and then (1) and (3) yield that there is an immersion f:Mm2m1f\colon M^{m}\looparrowright\mathbb{R}^{2m-1} such that 2[f]2[f] is 0 in Immint(m,m1)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(m,m-1). But then choosing any immersion g:Nn2ng\colon N^{n}\looparrowright\mathbb{R}^{2n} we get an immersion

f×g:M×N2k+1f\times g\colon M\times N\looparrowright\mathbb{R}^{2k+1}

which represents an element of order 22 in Immint(k+1,k)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(k+1,k) since we have 2[f×g]=2[f]×[g]=02[f\times g]=2[f]\times[g]=0 and Immint(,)\textstyle{\mathop{\rm Imm}}^{\mathrm{int}}(*,*) is a bi-graded ring. This finishes our proof.  \square

10 Morin maps

Recall that a stable map f:MnPn+kf\colon M^{n}\to P^{n+k} induces a stratification of MM according to its Thom–Boardman types (see [boar]): Σi(f)M\Sigma^{i}(f)\subset M denotes the submanifold consisting of the points where the rank of derivative of ff drops by ii, then we have Σi,j(f):=Σj(f|Σi(f))\Sigma^{i,j}(f):=\Sigma^{j}(f|_{\Sigma^{i}(f)}) and so on; this defines Σi1,,im(f)\Sigma^{i_{1},\ldots,i_{m}}(f) for any decreasing sequence i1,,imi_{1},\ldots,i_{m}. In the present section we shall consider maps which only have singularities of type Σ1\Sigma^{1}, 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 Σ1r:=Σ1,,1,0\Sigma^{1_{r}}:=\Sigma^{1,\ldots,1,0} (where the number of 11’s is rr) for some rr and each type Σ1r\Sigma^{1_{r}} contains precisely one singularity class; see [mor]. We shall denote the singularity class in Σ1r\Sigma^{1_{r}} also by the symbol Σ1r\Sigma^{1_{r}} and note that for r<sr<s we have Σ1r<Σ1s\Sigma^{1_{r}}<\Sigma^{1_{s}}, that is, Morin singularities form a single increasing sequence. We call Morin maps with at most Σ1r\Sigma^{1_{r}} singularity Σ1r\Sigma^{1_{r}}-maps and put

Morrσ(n,k):=Cob{Σ0,Σ11,,Σ1r}σ(n+k)\textstyle{\mathop{\rm Mor}}_{r}^{\sigma}(n,k):=\textstyle{\mathop{\rm Cob}}_{\{\Sigma^{0},\Sigma^{1_{1}},\ldots,\Sigma^{1_{r}}\}}^{\sigma}(\mathbb{R}^{n+k})

for any stable normal structure σ\sigma (except that for σ=Gξ\sigma=G\oplus\xi we decrease nn and increase kk by the rank of ξ\xi). Here r=r=\infty is also allowed and it means that we put no further restriction on Morin maps.

The rational cohomology of the Kazarian space KrSOK_{r}^{\mathrm{SO}} of oriented Σ1r\Sigma^{1_{r}}-maps (for 1r1\leq r\leq\infty) was computed by Szűcs [hosszu] and we can use this to obtain the form of φn+k\varphi_{n+k} rationally:

Proposition 10.1.

The endomorphism φn+k\varphi_{n+k}\otimes\mathbb{Q} of MorrSO(n,k)\textstyle{\mathop{\rm Mor}}_{r}^{\mathrm{SO}}(n,k)\otimes\mathbb{Q} is the following:

  1. (1)

    if k=2mk=2m is even, r<r<\infty is also even and qq is the number of non-negative integers a0,a1,,ama_{0},a_{1},\ldots,a_{m} such that a0a_{0} is odd and n=a0k(r+1)+𝑚i=1ai4in=a_{0}k(r+1)+\mathop{\overset{m}{\sum}}\limits_{\vbox to-1.50694pt{\kern-4.82224pt\hbox{$\scriptstyle i=1$}\vss}}a_{i}4i, then φn+k\varphi_{n+k}\otimes\mathbb{Q} is trivial on qq generators and the multiplication by 22 on the rest of the generators (in an appropriate basis),

  2. (2)

    otherwise φn+k\varphi_{n+k}\otimes\mathbb{Q} is the multiplication by 22.

Proof.  Put A:=[p1,,pm]A:=\mathbb{Q}[p_{1},\ldots,p_{m}] with degpi=4i\deg p_{i}=4i. By [hosszu, theorems 6 and 7] we have

MorrSO(,k)\displaystyle\textstyle{\mathop{\rm Mor}}_{r}^{\mathrm{SO}}(*,k)\otimes\mathbb{Q} H(KrSO;)=\displaystyle\cong H^{*}(K^{\mathrm{SO}}_{r};\mathbb{Q})=
={A,if k=2m is even and r is odd or A[er+1]/((er+1)2pmr+1),if k=2m and r are both evenand deger+1=k(r+1)A,if k=2m1 is odd and r=A/(pmr+12),if k=2m1 is odd and r<\displaystyle=\begin{cases}A,&\text{if }k=2m\text{ is even and }r\text{ is odd or }\infty\\ A[e^{r+1}]/\bigl{(}(e^{r+1})^{2}-p_{m}^{r+1}\bigr{)},&\text{if }k=2m\text{ and }r\text{ are both even}\\ &\text{and }\deg e^{r+1}=k(r+1)\\ A,&\text{if }k=2m-1\text{ is odd and }r=\infty\\ A/\Bigl{(}p_{m}^{\lceil\frac{r+1}{2}\rceil}\Bigr{)},&\text{if }k=2m-1\text{ is odd and }r<\infty\end{cases}

with piH4i(KrSO;)p_{i}\in H^{4i}(K^{\mathrm{SO}}_{r};\mathbb{Q}) and er+1Hk(r+1)(KrSO;)e^{r+1}\in H^{k(r+1)}(K^{\mathrm{SO}}_{r};\mathbb{Q}). Now following the action of the involution inducing the ι\iota of proposition 5.3 through the spectral sequences computed in [hosszu] yields that ι\iota changes the pip_{i} and er+1e^{r+1} here analogously to how it did in the proof of proposition 9.1, that is, we get pipip_{i}\mapsto p_{i} and er+1er+1e^{r+1}\mapsto-e^{r+1}. This implies our claim.  \square

As in the case of immersions, we only have an understanding of the torsion parts of these cobordism groups if the codimension kk is either small or relatively large. In particular we will investigate the cases when either we have k=1k=1 and 1r<1\leq r<\infty or kk is large compared to nn and rr is 11.

1-codimensional Morin maps

Morin maps of codimension 11 were considered by Szűcs [nulladik] who computed MorrSO(n,1)\textstyle{\mathop{\rm Mor}}_{r}^{\mathrm{SO}}(n,1) for 1r<1\leq r<\infty modulo small torsion groups. We also note that MorrO(n,1)\textstyle{\mathop{\rm Mor}}_{r}^{\mathrm{O}}(n,1) is finite 22-primary for all nn and 1r1\leq r\leq\infty by [szszt, theorem 1]. In the following we denote by 𝒞2\mathscr{C}_{2} the Serre class of finite 22-primary Abelian groups.

Proposition 10.2.

The endomorphism φn+1\varphi_{n+1} of MorrSO(n,1)\textstyle{\mathop{\rm Mor}}_{r}^{\mathrm{SO}}(n,1) is the multiplication by 22 modulo 𝒞2\mathscr{C}_{2} (that is, on the odd torsion and free parts) for 1r<1\leq r<\infty.

Proof.  We proceed by induction on rr. Observe that setting r=0r=0 we get the case of 11-codimensional immersions where we saw this claim to be true in the previous section. Now assume it to be true for r1r-1 and prove for rr.

We use the key fibration [hosszu, section 16] to obtain an exact sequence

Morr1SO(n,1)MorrSO(n,1)Immξ~rSO(n2r,2r+1)\textstyle{\mathop{\rm Mor}}_{r-1}^{\mathrm{SO}}(n,1)\to\textstyle{\mathop{\rm Mor}}_{r}^{\mathrm{SO}}(n,1)\to\textstyle{\mathop{\rm Imm}}^{\tilde{\xi}^{\mathrm{SO}}_{r}}(n-2r,2r+1)

where the first arrow is the natural forgetful homomorphism and the second one assigns to a cobordism class [f][f] the cobordism class of its restriction f|Σ1r(f)f|_{\Sigma^{1_{r}}(f)} to the most complicated stratum which is an immersion with normal bundle induced from ξ~rSO\tilde{\xi}^{\mathrm{SO}}_{r} (the universal normal bundle of the Σ1r\Sigma^{1_{r}}-stratum of oriented maps). These both commute with the reflection to a hyperplane (without changing orientation), i.e. the involution ι\iota in theorem 2.

By the considerations in [nulladik, section 2.2] for rr odd we have Immξ~rSO(n2r,2r+1)𝒞2\textstyle{\mathop{\rm Imm}}^{\tilde{\xi}^{\mathrm{SO}}_{r}}(n-2r,2r+1)\in\mathscr{C}_{2} and for rr even the forgetful homomorphism

πs(n2r)Immfr(n2r,2r+1)Immξ~rSO(n2r,2r+1)\pi^{s}(n-2r)\cong\textstyle{\mathop{\rm Imm}}^{\mathrm{fr}}(n-2r,2r+1)\to\textstyle{\mathop{\rm Imm}}^{\tilde{\xi}^{\mathrm{SO}}_{r}}(n-2r,2r+1)

(which assigns to the cobordism class of a framed immersion its class as an immersion with normal bundle induced from ξ~rSO\tilde{\xi}^{\mathrm{SO}}_{r}) is a 𝒞2\mathscr{C}_{2}-isomorphism. Observe that this homomorphism also commutes with the action of ι\iota.

In the previous section we saw that theorem 2 for 11-codimensional immersions is just the long exact sequence induced by the cofibration S1 2S1P2S^{1}\mathop{\xrightarrow{\leavevmode\nobreak\ 2\leavevmode\nobreak\ }}S^{1}\to\mathbb{R}P^{2} (where 22 denotes the degree-22 map), hence in that case the endomorphism φn2r+1\varphi_{n-2r+1} of ImmSO(n2r,1)πs(n2r)Immfr(n2r,2r+1)\textstyle{\mathop{\rm Imm}}^{\mathrm{SO}}(n-2r,1)\cong\pi^{s}(n-2r)\cong\textstyle{\mathop{\rm Imm}}^{\mathrm{fr}}(n-2r,2r+1) is 2id2\mathop{\rm id} which means that ι\iota acts identically on Immfr(n2r,2r+1)\textstyle{\mathop{\rm Imm}}^{\mathrm{fr}}(n-2r,2r+1). By the induction hypothesis ι\iota is the identity modulo 𝒞2\mathscr{C}_{2} on Morr1SO(n,1)\textstyle{\mathop{\rm Mor}}_{r-1}^{\mathrm{SO}}(n,1) too, thus it acts identically also on Immξ~rSO(n2r,2r+1)\textstyle{\mathop{\rm Imm}}^{\tilde{\xi}^{\mathrm{SO}}_{r}}(n-2r,2r+1) modulo 𝒞2\mathscr{C}_{2}. But then the exact sequence above shows that ι\iota has to be the identity modulo 𝒞2\mathscr{C}_{2} on MorrSO(n,1)\textstyle{\mathop{\rm Mor}}_{r}^{\mathrm{SO}}(n,1) as well which means that φn+1\varphi_{n+1} is 2id2\mathop{\rm id} modulo 𝒞2\mathscr{C}_{2}.  \square

Remark 10.3.

Using [nulladik, theorem A], a similar argument as in the proof above also shows that for r=1r=1 we have φn+1=2id\varphi_{n+1}=2\mathop{\rm id} not only modulo 𝒞2\mathscr{C}_{2} but even on the 22-primary part.

Corollary 10.4.

For all nn and 1r<1\leq r<\infty the group Morrint(n,1)\textstyle{\mathop{\rm Mor}}_{r}^{\mathrm{int}}(n,1) is finite 22-primary.

Proof.  This is immediate from the proposition above since φn+1\varphi_{n+1} is an isomorphism on the odd torsion part and the kernel and cokernel of φn+1\varphi_{n+1} on the free part are 0 and 2\mathbb{Z}_{2} respectively.  \square

This describes Morrint(n,1)\textstyle{\mathop{\rm Mor}}_{r}^{\mathrm{int}}(n,1) generally. Let us now consider the case r=1r=1, i.e. the so-called fold maps and put Foldint(n,1):=Mor1int(n,1)\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(n,1):=\textstyle{\mathop{\rm Mor}}_{1}^{\mathrm{int}}(n,1). Restricting the exact sequence of theorem 2 for fold cobordisms to the 22-primary parts [nulladik, theorem A] yields

ker(λ)n1 2idker(λ)n1Foldint(n,1)ker(λ)n2 2idker(λ)n2\ker(\lambda_{*})_{n-1}\mathop{\xrightarrow{\leavevmode\nobreak\ 2\mathop{\rm id}\leavevmode\nobreak\ }}\ker(\lambda_{*})_{n-1}\to\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(n,1)\to\ker(\lambda_{*})_{n-2}\mathop{\xrightarrow{\leavevmode\nobreak\ 2\mathop{\rm id}\leavevmode\nobreak\ }}\ker(\lambda_{*})_{n-2}

where (λ)m(\lambda_{*})_{m} denotes the Kahn–Priddy homomorphism πms(P)πs(m)\pi^{s}_{m}(\mathbb{R}P^{\infty})\to\pi^{s}(m) which maps onto the 22-primary torsion part of πs(m)\pi^{s}(m) (see [kp]).

Recalling the first few stable homotopy groups of P\mathbb{R}P^{\infty} computed by Liulevicius [liu] we obtain

mm 0 11 22 33 44 55 66 77 88 99
πms(P)\pi^{s}_{m}(\mathbb{R}P^{\infty}) 0 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} 8\mathbb{Z}_{8} 2\mathbb{Z}_{2} 0 2\mathbb{Z}_{2} 162\mathbb{Z}_{16}\oplus\mathbb{Z}_{2} (2)3(\mathbb{Z}_{2})^{3} (2)4(\mathbb{Z}_{2})^{4}
πs(m)\pi^{s}(m) \mathbb{Z} 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} 83\mathbb{Z}_{8}\oplus\mathbb{Z}_{3} 0 0 2\mathbb{Z}_{2} 1635\mathbb{Z}_{16}\oplus\mathbb{Z}_{3}\oplus\mathbb{Z}_{5} (2)2(\mathbb{Z}_{2})^{2} (2)3(\mathbb{Z}_{2})^{3}
ker(λ)m\ker(\lambda_{*})_{m} 0 0 0 0 2\mathbb{Z}_{2} 0 0 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} 2\mathbb{Z}_{2}

using the notation (2)r:=22(\mathbb{Z}_{2})^{r}:=\mathbb{Z}_{2}\oplus\ldots\oplus\mathbb{Z}_{2} with rr summands. This implies the following:

Proposition 10.5.

The cobordism groups Foldint(n,1)\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(n,1) for n10n\leq 10 are

nn 0 11 22 33 44 55 66 77 88 99 1010
Foldint(n,1)\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(n,1) 2\mathbb{Z}_{2} 0 0 0 0 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} 0 2\mathbb{Z}_{2} 2?2\mathbb{Z}_{2}\leavevmode\nobreak\ ?\leavevmode\nobreak\ \mathbb{Z}_{2} 2?2\mathbb{Z}_{2}\leavevmode\nobreak\ ?\leavevmode\nobreak\ \mathbb{Z}_{2}

where G?HG\leavevmode\nobreak\ ?\leavevmode\nobreak\ H again denotes the existence of a short exact sequence 0GG?HH00\to G\to G\leavevmode\nobreak\ ?\leavevmode\nobreak\ H\to H\to 0.

Large codimensional fold maps

In the following we shall consider fold maps of codimension kk such that the dimension of the source manifolds is 2k+12k+1 or 2k+22k+2 and, as we also did above, use the notation

Foldσ(n,k):=Mor1σ(n,k)\textstyle{\mathop{\rm Fold}}^{\sigma}(n,k):=\textstyle{\mathop{\rm Mor}}_{1}^{\sigma}(n,k)

for any stable normal structure σ\sigma (except that for σ=Gξ\sigma=G\oplus\xi we decrease nn and increase kk by the rank of ξ\xi). The cobordism groups FoldSO(n,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(n,k) and FoldO(n,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{O}}(n,k) 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 n<2k+1n<2k+1: if n<2k+1n<2k+1, then generic maps of nn-manifolds and (n+1)(n+1)-manifolds with codimension kk are stable and a computation with jet bundles implies that for any such map the codimension of the Σi\Sigma^{i}-stratum and the Σ1i\Sigma^{1_{i}}-stratum in the source manifold are i(k+i)i(k+i) and i(k+1)i(k+1) respectively, hence for n<2k+1n<2k+1 a generic kk-codimensional map of a cobordism of nn-manifolds is fold. Moreover, n=2k+1n=2k+1 and n=2k+2n=2k+2 are precisely the dimensions where fold cobordisms are not generic (that is, Foldσ(n,k)\textstyle{\mathop{\rm Fold}}^{\sigma}(n,k) 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. Σ1,1,0\Sigma^{1,1,0}.

As for the case of immersions, we will use the natural forgetful homomorphisms

αfoldSO:FoldSO(n,k)Ωn,αfoldint:Foldint(n,k)𝔚n,\displaystyle\alpha^{\mathrm{SO}}_{\mathrm{fold}}\colon\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(n,k)\to\Omega_{n},\quad\alpha^{\mathrm{int}}_{\mathrm{fold}}\colon\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(n,k)\to\mathfrak{W}_{n},
αfoldO:FoldO(n,k)𝔑nandαfoldO2detγ:FoldO2detγ(n2,k+2)𝔑n2\displaystyle\alpha^{\mathrm{O}}_{\mathrm{fold}}\colon\textstyle{\mathop{\rm Fold}}^{\mathrm{O}}(n,k)\to\mathfrak{N}_{n}\quad\text{and}\quad\alpha^{\mathrm{O}\oplus 2\det\gamma}_{\mathrm{fold}}\colon\textstyle{\mathop{\rm Fold}}^{\mathrm{O}\oplus 2\det\gamma}(n-2,k+2)\to\mathfrak{N}_{n-2}

that assign to the cobordism class of a fold map f:Mnn+kf\colon M^{n}\to\mathbb{R}^{n+k} (or a germ f~:ξMn+k\tilde{f}\colon\xi_{M}\to\mathbb{R}^{n+k}) with the appropriate normal structure the abstract cobordism class of MM. 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

\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FoldSO(n,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldSO\scriptstyle{\alpha^{\mathrm{SO}}_{\mathrm{fold}}}FoldSO(n,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldSO\scriptstyle{\alpha^{\mathrm{SO}}_{\mathrm{fold}}}Foldint(n,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}FoldSO(n1,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(n-1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldSO\scriptstyle{\alpha^{\mathrm{SO}}_{\mathrm{fold}}}\textstyle{\ldots}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωn\textstyle{\Omega_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2id\scriptstyle{2\mathop{\rm id}}Ωn\textstyle{\Omega_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚n\textstyle{\mathfrak{W}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωn1\textstyle{\Omega_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2id\scriptstyle{2\mathop{\rm id}}\textstyle{\ldots}

and

\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Foldint(n,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}FoldO(n,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{O}}(n,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldO\scriptstyle{\alpha^{\mathrm{O}}_{\mathrm{fold}}}FoldO2detγ(n2,k+2)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{O}\oplus 2\det\gamma}(n-2,k+2)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldO2detγ\scriptstyle{\alpha^{\mathrm{O}\oplus 2\det\gamma}_{\mathrm{fold}}}Foldint(n1,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(n-1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}\textstyle{\ldots}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\scriptstyle{0}𝔚n\textstyle{\mathfrak{W}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔑n\textstyle{\mathfrak{N}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔑n2\textstyle{\mathfrak{N}_{n-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\scriptstyle{0}𝔚n1\textstyle{\mathfrak{W}_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots}

where the rows are exact.

Lemma 10.7.

The map αfoldO2detγ\alpha^{\mathrm{O}\oplus 2\det\gamma}_{\mathrm{fold}} is an isomorphism FoldO2detγ(n2,k+2)𝔑n2\textstyle{\mathop{\rm Fold}}^{\mathrm{O}\oplus 2\det\gamma}(n-2,k+2)\cong\mathfrak{N}_{n-2} for all n2k+2n\leq 2k+2.

Proof.  This is almost completely analogous to lemma 9.7 by changing immersions to fold maps and k+1k+1 to 2k+22k+2. The only thing we have to add to its proof now is that the generic map

F:ξWn+k×[0,1]F\colon\xi_{W}\to\mathbb{R}^{n+k}\times[0,1]

(where WW is a cobordism of (n2)(n-2)-manifolds and ξW:=2detTW\xi_{W}:=2\det TW) only has fold singularities in a neighbourhood of the zero-section WW and its germ along WW satisfies the condition in definition 6.1 that the differential dF~d\tilde{F} restricted to any fibre of ξW\xi_{W} is injective.

We may assume that the singularity strata intersect WξWW\subset\xi_{W} transversally which firstly means that WW is disjoint from the cusp stratum (since the cusp stratum is at most 11-dimensional and WW is 22-codimensional in ξW\xi_{W}), hence it is indeed a fold map on a small neighbourhood of WW. Secondly it means that the local trivialisations of ξW\xi_{W} in a sufficiently small neighbourhood of the zero-section can be chosen such that for all pΣ1,0(F)p\in\Sigma^{1,0}(F) the whole fibre (in this neighbourhood) of ξW\xi_{W} containing pp belongs to Σ1,0(F)\Sigma^{1,0}(F), that is, both the fibres of ξW\xi_{W} and the fold stratum are orthogonal to WW.

Now for any point pΣ1,0(F)Wp\in\Sigma^{1,0}(F)\cap W a coordinate neighbourhood of pp has the form k+1×nk2×2\mathbb{R}^{k+1}\times\mathbb{R}^{n-k-2}\times\mathbb{R}^{2} where k+1×nk2\mathbb{R}^{k+1}\times\mathbb{R}^{n-k-2} and nk2×2\mathbb{R}^{n-k-2}\times\mathbb{R}^{2} are coordinate neighbourhoods in WW and in Σ1,0(F)\Sigma^{1,0}(F) respectively. By [rsz, theorem 6] the normal bundle of Σ1,0(F)\Sigma^{1,0}(F) is induced from γ1Oγ1OγkO\gamma_{1}^{\mathrm{O}}\oplus\gamma_{1}^{\mathrm{O}}\otimes\gamma_{k}^{\mathrm{O}} over BG1SOBG_{1}^{\mathrm{SO}} with G1SO={(ε,A)O(1)×O(k)εdetA=1}G_{1}^{\mathrm{SO}}=\{(\varepsilon,A)\in\mathrm{O}(1)\times\mathrm{O}(k)\mid\varepsilon\det A=1\}. The fibres of this in the coordinate neighbourhood of pp can be assumed to be tangent to k+1×{x}\mathbb{R}^{k+1}\times\{x\} for all xnk2×2x\in\mathbb{R}^{n-k-2}\times\mathbb{R}^{2} and can also be identified over all points xx. Thus in this neighbourhood the map FF has the desired form described in definition 6.1. If pp was instead in WΣ1,0(F)W\setminus\Sigma^{1,0}(F) then FF is an immersion near pp, hence then we also get the desired form of FF and so the first condition in definition 6.1 is always satisfied.  \square

Proposition 10.8.

For all k1k\geq 1 we have

  1. (1)

    Foldint(2k+1,k)𝔚2k+1\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(2k+1,k)\cong\mathfrak{W}_{2k+1} if k2k\neq 2,

  2. (2)

    Foldint(5,2)𝔚5222\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(5,2)\cong\mathfrak{W}_{5}\oplus\mathbb{Z}_{2}\cong\mathbb{Z}_{2}\oplus\mathbb{Z}_{2}.

Proof.  We shall use the first diagram above lemma 10.7 and obtain FoldSO(2k+1,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+1,k) from [eszt, theorem 1].

Proof of (1).  If k>2k>2 is even, then αfoldSO\alpha^{\mathrm{SO}}_{\mathrm{fold}} is an isomorphism in this dimension yielding the diagram

Ω2k+1\textstyle{\Omega_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}2id\scriptstyle{2\mathop{\rm id}}Ω2k+1\textstyle{\Omega_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}Foldint(2k+1,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(2k+1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}Ω2k\textstyle{\Omega_{2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}2id\scriptstyle{2\mathop{\rm id}}Ω2k\textstyle{\Omega_{2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}Ω2k+1\textstyle{\Omega_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2id\scriptstyle{2\mathop{\rm id}}Ω2k+1\textstyle{\Omega_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚2k+1\textstyle{\mathfrak{W}_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω2k\textstyle{\Omega_{2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2id\scriptstyle{2\mathop{\rm id}}Ω2k\textstyle{\Omega_{2k}}

and now the 55-lemma implies the isomorphism claimed.

For kk odd we have FoldSO(2k+1,k)Ω2k+13t\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+1,k)\cong\Omega_{2k+1}\oplus\mathbb{Z}_{3^{t}} (with an appropriate number tt) and this gives

Ω2k+13t\textstyle{\Omega_{2k+1}\oplus\mathbb{Z}_{3^{t}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id+ι\scriptstyle{\mathop{\rm id}+\iota}Ω2k+13t\textstyle{\Omega_{2k+1}\oplus\mathbb{Z}_{3^{t}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}Foldint(2k+1,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(2k+1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}Ω2k\textstyle{\Omega_{2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω2k+1\textstyle{\Omega_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚2k+1\textstyle{\mathfrak{W}_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω2k\textstyle{\Omega_{2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

thus kerαfoldint\ker\alpha^{\mathrm{int}}_{\mathrm{fold}} is 2u\mathbb{Z}_{2^{u}} for some 0ut0\leq u\leq t depending on how ι\iota acts on the 3t\mathbb{Z}_{3^{t}} part of FoldSO(2k+1,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+1,k). Now applying lemma 10.7, part of the second diagram above it has the form

𝔑2k\textstyle{\mathfrak{N}_{2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ3k+2\scriptstyle{\psi^{\prime}_{3k+2}}Foldint(2k+1,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(2k+1,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}φ3k+1\scriptstyle{\varphi^{\prime}_{3k+1}}𝔑2k+1\textstyle{\mathfrak{N}_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚2k+1\textstyle{\mathfrak{W}_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔑2k+1\textstyle{\mathfrak{N}_{2k+1}}

hence we have 3ukerαfoldint=kerφ3k+1=imψ3k+2\mathbb{Z}_{3^{u}}\cong\ker\alpha^{\mathrm{int}}_{\mathrm{fold}}=\ker\varphi^{\prime}_{3k+1}=\mathop{\rm im}\psi^{\prime}_{3k+2} which is isomorphic to a factor group of 𝔑2k\mathfrak{N}_{2k}. But 𝔑2k\mathfrak{N}_{2k} is 22-primary so this can only happen if u=0u=0, that is, if αfoldint\alpha^{\mathrm{int}}_{\mathrm{fold}} is an isomorphism.

Proof of (2).  We have FoldSO(4,2)Ω4\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(4,2)\cong\Omega_{4}\cong\mathbb{Z} and FoldSO(5,2)Ω5222\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(5,2)\cong\Omega_{5}\oplus\mathbb{Z}_{2}\cong\mathbb{Z}_{2}\oplus\mathbb{Z}_{2} and their endomorphisms φ6\varphi_{6} and φ7\varphi_{7} are both the multiplication by 22, hence we get the diagram

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω52\textstyle{\Omega_{5}\oplus\mathbb{Z}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}Foldint(5,2)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(5,2)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω5\textstyle{\Omega_{5}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚5\textstyle{\mathfrak{W}_{5}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

which proves our statement.  \square

Proposition 10.9.

For all k1k\geq 1 we have

  1. (1)

    Foldint(2k+2,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(2k+2,k) is isomorphic to a subgroup of 𝔚2k+2\mathfrak{W}_{2k+2} of index 22 if k2k\neq 2,

  2. (2)

    Foldint(6,2)𝔚6222\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(6,2)\cong\mathfrak{W}_{6}\oplus\mathbb{Z}_{2}\cong\mathbb{Z}_{2}\oplus\mathbb{Z}_{2}.

Proof.  We will again always use the first diagram above lemma 10.7 and we know FoldSO(2k+1,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+1,k) and FoldSO(2k+2,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+2,k) from [eszt, theorem 1] and [2k+2, theorem 4] respectively.

Proof of (1).  If k>2k>2 is even, then αfoldSO\alpha^{\mathrm{SO}}_{\mathrm{fold}} is an isomorphism FoldSO(2k+1,k)Ω2k+1\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+1,k)\cong\Omega_{2k+1} and the embedding of FoldSO(2k+2,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+2,k) into Ω2k+2\Omega_{2k+2} as an index-22 subgroup.

If kk is odd, then we have FoldSO(2k+1,k)Ω2k+13t\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+1,k)\cong\Omega_{2k+1}\oplus\mathbb{Z}_{3^{t}} but, as we saw in the proof of proposition 10.8, the 3t\mathbb{Z}_{3^{t}} part is mapped isomorphically by the endomorphism φ3k+1\varphi_{3k+1}, hence it does not appear in its kernel. Moreover, for kk odd αfoldSO\alpha^{\mathrm{SO}}_{\mathrm{fold}} embeds FoldSO(2k+2,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+2,k) into Ω2k+2\Omega_{2k+2} as the kernel of the homomorphism p¯k+12[]:Ω2k+2\overline{p}_{\frac{k+1}{2}}[\cdot]\colon\Omega_{2k+2}\to\mathbb{Z} which maps a cobordism class [M][M] to the normal Pontryagin number p¯k+12[M]\overline{p}_{\frac{k+1}{2}}[M]. Now if we take the cokernel of φ3k+2\varphi_{3k+2}, that is, we factor kerp¯k+12[]\ker\overline{p}_{\frac{k+1}{2}}[\cdot] by 2kerp¯k+12[]2\ker\overline{p}_{\frac{k+1}{2}}[\cdot] and Ω2k+2\Omega_{2k+2} by 2Ω2k+22\Omega_{2k+2}, then the image of FoldSO(2k+2,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(2k+2,k) under this quotient map will again become an index-22 subgroup.

Hence both for kk odd and for kk even we obtain that there is a group GG such that the commutative diagram

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G\textstyle{G\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Foldint(2k+2,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(2k+2,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}Ω2k+1\textstyle{\Omega_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω2k+22Ω2k+2\textstyle{\frac{\Omega_{2k+2}}{2\Omega_{2k+2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚2k+2\textstyle{\mathfrak{W}_{2k+2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω2k+1\textstyle{\Omega_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

holds where the monomorphism from GG into Ω2k+2/2Ω2k+2\Omega_{2k+2}/2\Omega_{2k+2} has cokernel 2\mathbb{Z}_{2}. This implies our statement.

Proof of (2).  We have FoldSO(5,2)Ω5222\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(5,2)\cong\Omega_{5}\oplus\mathbb{Z}_{2}\cong\mathbb{Z}_{2}\oplus\mathbb{Z}_{2} and FoldSO(6,2)Ω60\textstyle{\mathop{\rm Fold}}^{\mathrm{SO}}(6,2)\cong\Omega_{6}\cong 0 which gives us the diagram

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Foldint(6,2)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(6,2)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}Ω52\textstyle{\Omega_{5}\oplus\mathbb{Z}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id0\scriptstyle{\mathop{\rm id}\oplus 0}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚6\textstyle{\mathfrak{W}_{6}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω5\textstyle{\Omega_{5}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

proving our claim.  \square

Corollary 10.10.

For all k2k\neq 2 the homomorphism αfoldO\alpha^{\mathrm{O}}_{\mathrm{fold}} embeds FoldO(2k+2,k)\textstyle{\mathop{\rm Fold}}^{\mathrm{O}}(2k+2,k) into 𝔑2k+2\mathfrak{N}_{2k+2} as a subgroup of index 22.

Proof.  We apply lemma 10.7, the second diagram above it and also that the homomorphism ψ3k+3:FoldO2detγ(2k+1,k+2)Foldint(2k+2,k)\psi^{\prime}_{3k+3}\colon\textstyle{\mathop{\rm Fold}}^{\mathrm{O}\oplus 2\det\gamma}(2k+1,k+2)\to\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(2k+2,k) is zero as we saw in the proof of proposition 10.8. Then we have

Foldint(2k+2,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{int}}(2k+2,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldint\scriptstyle{\alpha^{\mathrm{int}}_{\mathrm{fold}}}FoldO(2k+2,k)\textstyle{\textstyle{\mathop{\rm Fold}}^{\mathrm{O}}(2k+2,k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αfoldO\scriptstyle{\alpha^{\mathrm{O}}_{\mathrm{fold}}}𝔑2k\textstyle{\mathfrak{N}_{2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔚2k+2\textstyle{\mathfrak{W}_{2k+2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔑2k+2\textstyle{\mathfrak{N}_{2k+2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔑2k\textstyle{\mathfrak{N}_{2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

thus the proposition above implies that αfoldO\alpha^{\mathrm{O}}_{\mathrm{fold}} is a monomorphism with cokernel 2\mathbb{Z}_{2}.  \square

Remark 10.11.

This corollary contradicts [2k+2, theorem 4.a)] which essentially claims that both the kernel and the cokernel of

αfoldO:FoldO(2k+2,k)𝔑2k+2\alpha^{\mathrm{O}}_{\mathrm{fold}}\colon\textstyle{\mathop{\rm Fold}}^{\mathrm{O}}(2k+2,k)\to\mathfrak{N}_{2k+2}

are 2\mathbb{Z}_{2}. The problem with this, according to Terpai, is the following: the computations in [2k+2] show that here kerαfoldO\ker\alpha^{\mathrm{O}}_{\mathrm{fold}} appears as the factor by 2\mathbb{Z}_{2} of the group 22\mathbb{Z}_{2}\oplus\mathbb{Z}_{2} of a complete set of invariants of cusp null-cobordisms of fold maps (recall that a null-cobordism is a map of a (2k+3)(2k+3)-manifold with codimension kk and such a map generically only has Σ1,0\Sigma^{1,0} (fold) and Σ1,1,0\Sigma^{1,1,0} (cusp) singularities); one of the summands 2\mathbb{Z}_{2} (which gets cancelled by the factoring) measures the number of components of the cusp curve of such a null-cobordism modulo 22 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 2\mathbb{Z}_{2} does not appear as the orientation cannot change over this curve. Hence kerαfoldO\ker\alpha^{\mathrm{O}}_{\mathrm{fold}} is actually the factor of only 2\mathbb{Z}_{2} by 2\mathbb{Z}_{2}, that is, αfoldO\alpha^{\mathrm{O}}_{\mathrm{fold}} is mono. Moreover, it is also apparent that its image is the kernel of the Thom polynomial of cusp singularity (w¯k+12+w¯kw¯k+2)[]:𝔑2k+22(\overline{w}_{k+1}^{2}+\overline{w}_{k}\overline{w}_{k+2})[\cdot]\colon\mathfrak{N}_{2k+2}\to\mathbb{Z}_{2} which maps a cobordism class [M][M] to the normal Stiefel–Whitney number w¯k+12[M]+w¯kw¯k+2[M]\overline{w}_{k+1}^{2}[M]+\overline{w}_{k}\overline{w}_{k+2}[M].

11 Bordisms

In this final section we point out that theorems 2 and 2 also hold when we change τ\tau-cobordism groups to general bordism groups (although the set of all possible singularities of kk-codimensional germs generally contains unstable ones which we excluded from τ\tau). 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 QQ there is a long exact sequence

Ωn(Q)φnΩn(Q)χn𝔚n(Q)ψnΩn1(Q)\ldots\to\Omega_{n}(Q)\mathop{\xrightarrow{\leavevmode\nobreak\ \varphi_{n}\leavevmode\nobreak\ }}\Omega_{n}(Q)\mathop{\xrightarrow{\leavevmode\nobreak\ \chi_{n}\leavevmode\nobreak\ }}\mathfrak{W}_{n}(Q)\mathop{\xrightarrow{\leavevmode\nobreak\ \psi_{n}\leavevmode\nobreak\ }}\Omega_{n-1}(Q)\to\ldots

and a short exact sequence

0𝔚n(Q)φn𝔑n(Q)χn𝔑n2(Q)00\to\mathfrak{W}_{n}(Q)\mathop{\xrightarrow{\leavevmode\nobreak\ \varphi^{\prime}_{n}\leavevmode\nobreak\ }}\mathfrak{N}_{n}(Q)\mathop{\xrightarrow{\leavevmode\nobreak\ \chi^{\prime}_{n}\leavevmode\nobreak\ }}\mathfrak{N}_{n-2}(Q)\to 0

where φn\varphi_{n} is the multiplication by 22, χn\chi_{n} and φn\varphi^{\prime}_{n} are the forgetful homomorphisms, ψm\psi_{m} assigns to the bordism class of the map of a Wall manifold f:MQf\colon M\to Q the bordism class of f|PD(w1(M))f|_{\mathop{\rm PD}(w_{1}(M))} and χn\chi^{\prime}_{n} assigns to the bordism class of a map g:NQg\colon N\to Q the bordism class of g|PD(w1(N)2)g|_{\mathop{\rm PD}(w_{1}(N)^{2})}.

Proof.  The nn’th bordism groups of any CW complex are the same as those of its (n+1)(n+1)-skeleton, hence for the purpose of this proof we may only take a finite dimensional skeleton of QQ instead of the actual QQ. 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 QQ is an orientable manifold of dimension qq.

Fix a number nn and let rr be such that generic maps of (n+1)(n+1)-manifolds to rr-manifolds are embeddings. Then we have

Ωn(Q)CobτSO(Q×r),𝔚n(Q)Cobτint(Q×r)and𝔑n(Q)CobτO(Q×r)\Omega_{n}(Q)\cong\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{SO}}(Q\times\mathbb{R}^{r}),\quad\mathfrak{W}_{n}(Q)\cong\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{int}}(Q\times\mathbb{R}^{r})\quad\text{and}\quad\mathfrak{N}_{n}(Q)\cong\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}}(Q\times\mathbb{R}^{r})

for any set τ\tau of stable singularities of codimension q+rnq+r-n.

Claim.  𝔑n2(Q)CobτO2detγ(Q×r)\mathfrak{N}_{n-2}(Q)\cong\textstyle{\mathop{\rm Cob}}_{\tau}^{\mathrm{O}\oplus 2\det\gamma}(Q\times\mathbb{R}^{r}).

Proof.  Let f:Mn2Qf\colon M^{n-2}\to Q be any map of an (n2)(n-2)-manifold to QQ and put ξM:=2detνf\xi_{M}:=2\det\nu_{f}. Now let f~:ξMQ\tilde{f}\colon\xi_{M}\to Q be any map that extends the map ff given on the zero-section of ξM\xi_{M} (such an extension is unique up to homotopies fixed on the zero-section) and let i:ξMri\colon\xi_{M}\hookrightarrow\mathbb{R}^{r} be any embedding (which is also unique up to isotopy). Then assigning to t