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

Spaces of polynomials with constrained divisors as Grassmanians for traversing flows

Gabriel Katz MIT, Department of Mathematics, 77 Massachusetts Ave., Cambridge, MA 02139, U.S.A. [email protected]
(Date: 03/22/2022)
Abstract.

We study traversing vector flows vv on smooth compact manifolds XX with boundary. For a given compact manifold X^\hat{X}, equipped with a traversing vector field v^\hat{v} which is convex with respect to X^\partial\hat{X}, we consider submersions/embeddings α:XX^\alpha:X\to\hat{X} such that dimX=dimX^\dim X=\dim\hat{X} and α(X)\alpha(\partial X) avoids a priory chosen tangency patterns Θ\Theta to the v^\hat{v}-trajectories. In particular, for each v^\hat{v}-trajectory γ^\hat{\gamma}, we restrict the cardinality of γ^α(X)\hat{\gamma}\cap\alpha(\partial X) by an even number dd. We call (X^,v^)(\hat{X},\hat{v}) a convex pseudo-envelop/envelop of the pair (X,v)(X,v). Here the vector field v=α(v^)v=\alpha^{\dagger}(\hat{v}) is the α\alpha-transfer of v^\hat{v} to XX.

For a fixed (X^,v^)(\hat{X},\hat{v}), we introduce an equivalence relation among convex pseudo-envelops/ envelops α:(X,v)(X^,v^)\alpha:(X,v)\to(\hat{X},\hat{v}), which we call a quasitopy. The notion of quasitopy is a crossover between bordisms of pseudo-envelops and their pseudo-isotopies. In the study of quasitopies 𝒬𝒯d(Y,𝐜Θ)\mathcal{QT}_{d}(Y,\mathbf{c}\Theta), the spaces 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta} of real univariate polynomials of degree dd with real divisors whose combinatorial types avoid the closed poset Θ\Theta play the classical role of Grassmanians.

We compute, in the homotopy-theoretical terms that involve (X^,v^)(\hat{X},\hat{v}) and 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta}, the quasitopies of convex envelops which avoid the Θ\Theta-tangency patterns. We introduce characteristic classes of pseudo-envelops and show that they are invariants of their quasitopy classes. Then we prove that the quasitopies 𝒬𝒯d(Y,𝐜Θ)\mathcal{QT}_{d}(Y,\mathbf{c}\Theta) often stabilize, as dd\to\infty.

1. Introduction

This paper is the second in a series, which is inspired by the works of Arnold [Ar], [Ar1], and Vassiliev [V]. It is a direct continuation of [K9], where many ideas of this article are present. Here we apply these ideas to traversing flows on manifolds with boundary. As [K9], this article relies heavily on computations from [KSW1], [KSW2]. While [K9] is studying immersions of compact nn-dimensional manifolds into products ×Y\mathbb{R}\times Y, where YY is a compact nn-dimensional manifold, the present paper deals with special traversing flows on compact (n+1)(n+1)-dimensional manifolds XX, the flows that admit a “virtual global section”. The existence of such sections allows to establish a transparent correspondence between the universe of immersions into the products and the universe of traversing flows with virtual sections. In fact, with this correspondence in place, many results of the present article are just reformulations of the corresponding results from [K9], thus motivating and justifying them. However, we prove also many propositions with no analogues in [K9] (for example, Proposition 4.1, Theorem 4.1, Corollary 4.4, Proposition 4.2, and Theorem 4.8).

Let us describe here our results informally, in a manner that clarifies their nature, but does not involve their most general forms, which carry the burden of combinatorial decorations.

As we have mentioned above, we study traversing vector flows vv (see Definition 4.1 and [K1], [K2]) on smooth compact (n+1)(n+1)-dimensional manifolds XX with boundary. Every traversing flow admits a Lyapunov function, and this fact may serve as a working definition of such flows. We are interested in the combinatorial patterns ω\omega that describe how the vv-trajectories are tangent to the boundary X\partial X. Specifically, we are concerned with the traversing vector flows whose tangency patterns ω\omega do not belong to a given closed poset Θ\Theta. Depending on Θ\Theta, the very existence of such flows puts strong restrictions on the topology of XX (see [K1]-[K8], [K10]). For example, in Fig.1, Θ\Theta includes the cubic and the double tangencies of X\partial X to the family of uu-directed lines. In other words, in this figure, the tangency patterns ω=(3)\omega=(3) and ω=(22)\omega=(22) are forbidden.

In order to simplify the investigation of such flows, we “envelop” them into, so called, convex pseudo-envelops (X^,v^)(\hat{X},\hat{v}). A convex pseudo-envelop consists of: (i) a compact smooth (n+1)(n+1)-manifold X^\hat{X}, (ii) a traversing vector field v^\hat{v} on it such that the boundary X^\partial\hat{X} is convex (see Definition 4.5) with respect to the v^\hat{v}-flow, (iii) a submersion α:XX^\alpha:X\to\hat{X} such that α(X)\alpha(\partial X) avoids a forbidden tangency patterns Θ\Theta to the v^\hat{v}-trajectories, and (iv) the pull-back α(v^)\alpha^{\dagger}(\hat{v}) of v^\hat{v}, coinciding with the given vector field vv (see Fig.1, in which X^\hat{X} is a cylinder).

Refer to caption


Figure 1. Convex pseudo-envelops α:(X,α(u))(X^,u)\alpha:(X,\alpha^{\dagger}(\partial_{u}))\to(\hat{X},\partial_{u}) of a punctured torus XX (on the top), and of a punctured surface XX of genus 22 (on the bottom). The envelop X^\hat{X} is the cylinder [0,1]×S1[0,1]\times S^{1}, equipped with the constant vector field u\partial_{u}. Both submersions α\alpha are generic relative to the vertical vector field u\partial_{u} on the cylinder. The dots mark the multiplicity 22 tangencies of α(X)\alpha(\partial X) to the θ\theta-fibers; the self-intersections of α(X)\alpha(\partial X) are unmarked. In both examples, the cardinality of the fibers θα:XS1\theta\circ\alpha^{\partial}:\partial X\to S^{1} does not exceed 66.

The existence of a convex pseudo-envelop provides the vv-flow on XX with a “virtual global section”, a significantly effective tool. Its existence allows us to transfer the results from [K9] about immersions/embeddings {β:M×Y}\{\beta:M\to\mathbb{R}\times Y\}, where MM and YY are compact smooth nn-manifolds, to the parallel results about convex envelops of traversing flows.

Next, we organize convex pseudo-envelops {α:(X,v)(X^,v^)}\{\alpha:(X,v)\to(\hat{X},\hat{v})\} with the fixed target (X^,v^)(\hat{X},\hat{v}) into the quasitopy equivalence classes 𝒬𝒯d(X^,v^;𝐜Θ)\mathcal{QT}_{d}(\hat{X},\hat{v};\mathbf{c}\Theta), where dd is a given even natural number. For a typical v^\hat{v}-trajectory γ^X^\hat{\gamma}\subset\hat{X}, dd gives an upper bound of the cardinality of the set γ^α(X)\hat{\gamma}\cap\alpha(\partial X). Fig. 3 may give some impression of the nature of quasitopy; it depicts a pair of submersions α0,α1\alpha_{0},\alpha_{1}, linked by a kind of cobordism with a strict combinatorial control of its tangent patterns.

The quasitopies come in two flavors: 𝒬𝒯d𝖾𝗆𝖻(X^,v^;𝐜Θ)\mathcal{QT}_{d}^{\mathsf{emb}}(\hat{X},\hat{v};\mathbf{c}\Theta), formed by regular embeddings α\alpha, and 𝒬𝒯d𝗌𝗎𝖻(X^,v^;𝐜Θ)\mathcal{QT}_{d}^{\mathsf{sub}}(\hat{X},\hat{v};\mathbf{c}\Theta), formed by more general submersions α\alpha.

Consider the locus 1+X^(v^)X^\partial_{1}^{+}\hat{X}(\hat{v})\subset\partial\hat{X}, where v^\hat{v} points inside of X^\hat{X}, and the locus 2X^(v^)1+X^(v^)\partial_{2}^{-}\hat{X}(\hat{v})\subset\partial_{1}^{+}\hat{X}(\hat{v}), where v^\hat{v} is tangent to X^\partial\hat{X}. Then we construct two classifying maps

(1.1) Φ𝖾𝗆𝖻:𝒬𝒯d𝖾𝗆𝖻(X^,v^;𝐜Θ)[(1+X^(v^),2X^(v^)),(𝒫d𝐜Θ,pt)],\displaystyle\Phi^{\mathsf{emb}}:\mathcal{QT}_{d}^{\mathsf{emb}}(\hat{X},\hat{v};\mathbf{c}\Theta)\to[(\partial_{1}^{+}\hat{X}(\hat{v}),\partial_{2}^{-}\hat{X}(\hat{v})),\,(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt)],
Φ𝗌𝗎𝖻:𝒬𝒯d𝗌𝗎𝖻(X^,v^;𝐜Θ)[(1+X^(v^),2X^(v^)),(𝒫d𝐜Θ,pt)],\displaystyle\Phi^{\mathsf{sub}}:\mathcal{QT}_{d}^{\mathsf{sub}}(\hat{X},\hat{v};\mathbf{c}\Theta)\to[(\partial_{1}^{+}\hat{X}(\hat{v}),\partial_{2}^{-}\hat{X}(\hat{v})),\,(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt)],

whose targets are sets of homotopy classes from the quotient space 1+X^(v^)/2X^(v^)\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v}) to the universal space 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta}—the “Grassmanian”—, formed by real polynomials of degree dd whose real divisors avoid a given poset Θ\Theta.

Then we prove (see Theorem 4.2) that Φ𝖾𝗆𝖻\Phi^{\mathsf{emb}} is a bijection and Φ𝗌𝗎𝖻\Phi^{\mathsf{sub}} is a split surjection.

For a constant vector field v^\hat{v} on the standard ball Dn+1D^{n+1}, the sets 𝒬𝒯d𝗌𝗎𝖻/𝖾𝗆𝖻(Dn+1,v^;𝐜Θ)\mathcal{QT}_{d}^{\mathsf{sub/emb}}(D^{n+1},\hat{v};\mathbf{c}\Theta) are groups; the group operation is induced by the boundary connected sum of pseudo-envelops/envelops. These groups are abelian for n>1n>1. The group 𝒬𝒯d𝗌𝗎𝖻(Dn+1,v^;𝐜Θ)\mathcal{QT}_{d}^{\mathsf{sub}}(D^{n+1},\hat{v};\mathbf{c}\Theta) is a split extension of 𝒬𝒯d𝖾𝗆𝖻(Dn+1,v^;𝐜Θ)\mathcal{QT}_{d}^{\mathsf{emb}}(D^{n+1},\hat{v};\mathbf{c}\Theta) (see Theorem 4.4). We prove that

𝒬𝒯d𝖾𝗆𝖻(Dn+1,v^;𝐜Θ)πn(𝒫d𝐜Θ,pt).\mathcal{QT}_{d}^{\mathsf{emb}}(D^{n+1},\hat{v};\mathbf{c}\Theta)\approx\pi_{n}(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt).

In the process, in the Subsection 4.2, we notice an interesting relation between the connected components of the of the space of convex traversing flows on the standard (n+1)(n+1)-ball and the monoid 𝐇𝐒n1\mathbf{HS}_{n-1} of smooth types of integral homology (n1)(n-1)-spheres, being considered up to connected sums with smooth homotopy (n1)(n-1)-spheres (see Proposition 4.1).

The groups 𝒬𝒯d𝗌𝗎𝖻/𝖾𝗆𝖻(Dn+1,v^;𝐜Θ)\mathcal{QT}_{d}^{\mathsf{sub/emb}}(D^{n+1},\hat{v};\mathbf{c}\Theta) act on the quasitopies of convex pseudo-envelops 𝒬𝒯d𝗌𝗎𝖻/𝖾𝗆𝖻(X^,v^;𝐜Θ)\mathcal{QT}_{d}^{\mathsf{sub/emb}}(\hat{X},\hat{v};\mathbf{c}\Theta), and the classifying maps Φ𝗌𝗎𝖻/𝖾𝗆𝖻\Phi^{\mathsf{sub/emb}} in (1.1) are equivariant.

These classifying maps deliver a variety of (co)homotopy and (co)homology invariants of traversing flows which admit convex pseudo-envelops (see Theorem 4.7, Proposition 4.6, and Theorem 4.8).

We also establish several of results (Theorem 4.3, Corollary 4.6) about the stabilization of the sets 𝒬𝒯d𝖾𝗆𝖻(X^,v^;𝐜Θ)\mathcal{QT}_{d}^{\mathsf{emb}}(\hat{X},\hat{v};\mathbf{c}\Theta) as dd\to\infty.

Finally, we compute 𝒬𝒯d𝖾𝗆𝖻(X^,v^;𝐜Θ)\mathcal{QT}_{d}^{\mathsf{emb}}(\hat{X},\hat{v};\mathbf{c}\Theta) for may special cases of Θ\Theta and (X^,v^)(\hat{X},\hat{v}).

2. Spaces of real polynomials with constrained real divisors

For our reader convenience, we state here a number of results from [KSW1] and [KSW2] about the topology of spaces of real monic univariate polynomials with constrained real divisors. These results are crucial for the applications to follow. To make the present paper even partially self-contained, we are basically recycling Section 2 from [K9].

Let 𝒫d\mathcal{P}_{d} denote the space of real monic univariate polynomials of degree dd. Given a polynomial P(u)=ud+ad1ud1++a0P(u)=u^{d}+a_{d-1}u^{d-1}+\cdots+a_{0} with real coefficients, we consider its real divisor D(P)D_{\mathbb{R}}(P). Let ωi\omega_{i} denotes the multiplicity of the ii-th real root of PP, the real roots being ordered by their magnitude. We call the ordered \ell-tuple of natural numbers ω=(ω1,,ω)\omega=(\omega_{1},\ldots,\omega_{\ell}) the real root multiplicity pattern of P(u)P(u), or the multiplicity pattern for short.

Such sequences ω\omega form a universal poset (𝛀,)(\mathbf{\Omega},\succ). The partial order “\succ” in 𝛀\mathbf{\Omega} is defined in terms of two types of elementary operations: merges {𝖬i}i\{\mathsf{M}_{i}\}_{i} and inserts {𝖨i}i\{\mathsf{I}_{i}\}_{i} The operation 𝖬i\mathsf{M}_{i} merges a pair of adjacent entries ωi,ωi+1\omega_{i},\omega_{i+1} of ω=(ω1,,ωi,ωi+1,,ωq)\omega=(\omega_{1},\dots,\omega_{i},\omega_{i+1},\dots,\omega_{q}) into a single component ω~i=ωi+ωi+1\tilde{\omega}_{i}=\omega_{i}+\omega_{i+1}, thus forming a new shorter sequence 𝖬i(ω)=(ω1,,ω~i,,ωq)\mathsf{M}_{i}(\omega)=(\omega_{1},\dots,\tilde{\omega}_{i},\dots,\omega_{q}). The insert operation 𝖨i\mathsf{I}_{i} either insert 22 in-between ωi\omega_{i} and ωi+1\omega_{i+1}, thus forming a new longer sequence 𝖨i(ω)=(,ωi,2,ωi+1,)\mathsf{I}_{i}(\omega)=(\dots,\omega_{i},2,\omega_{i+1},\dots), or, in the case of 𝖨0\mathsf{I}_{0}, appends 22 before the sequence ω\omega, or, in the case 𝖨q\mathsf{I}_{q}, appends 22 after the sequence ω\omega.

We define the order ωω\omega\succ\omega^{\prime}, where ω,ω𝛀\omega,\omega^{\prime}\in\mathbf{\Omega}, if one can produce ω\omega^{\prime} from ω\omega by applying a sequence of these elementary operations.

For a sequence ω=(ω1,ω2,,ωq)𝛀\omega=(\omega_{1},\omega_{2},\,\dots\,,\omega_{q})\in\mathbf{\Omega}, we introduce the norm and the reduced norm of ω\omega by the formulas:

(2.1) |ω|=𝖽𝖾𝖿iωiand|ω|=𝖽𝖾𝖿i(ωi1).\displaystyle|\omega|=_{\mathsf{def}}\sum_{i}\;\omega_{i}\quad\text{and}\quad|\omega|^{\prime}=_{\mathsf{def}}\sum_{i}\;(\omega_{i}-1).

Note that qq, the cardinality of the support of ω\omega, is equal to |ω||ω||\omega|-|\omega|^{\prime}.

Let 𝖱̊dω\mathring{{\mathsf{R}}}^{\omega}_{d} be of the set of all polynomials with the real root multiplicity pattern ω\omega, and let 𝖱dω{\mathsf{R}}_{d}^{\omega} be its closure in 𝒫d\mathcal{P}_{d}.

For a given collection Θ\Theta of multiplicity patterns {ω}\{\omega\}, which share the pairity of their norms {|ω|}\{|\omega|\} and is closed under the merge and insert operations, we consider the union 𝒫dΘ\mathcal{P}_{d}^{\Theta} of the subspaces 𝖱̊dω\mathring{{\mathsf{R}}}_{d}^{\omega}, taken over all ωΘ\omega\in\Theta such that |ω|d|\omega|\leq d and |ω|dmod2|\omega|\equiv d\mod 2. We denote by 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta} its complement 𝒫d𝒫dΘ\mathcal{P}_{d}\setminus\mathcal{P}^{\Theta}_{d}.

Since 𝒫dΘ\mathcal{P}^{\Theta}_{d} is contractible, it makes more sense to consider its one-point compactification 𝒫¯dΘ\bar{\mathcal{P}}_{d}^{\Theta}. If the set 𝒫¯dΘ\bar{\mathcal{P}}_{d}^{\Theta} is closed in 𝒫¯d\bar{\mathcal{P}}_{d}, by the Alexander duality on the sphere 𝒫¯dSd\bar{\mathcal{P}}_{d}\cong S^{d}, we get

Hj(𝒫d𝐜Θ;)Hdj1(𝒫¯dΘ;).H^{j}(\mathcal{P}_{d}^{\mathbf{c}\Theta};\mathbb{Z})\approx H_{d-j-1}(\bar{\mathcal{P}}_{d}^{\Theta};\mathbb{Z}).

This implies that the spaces 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta} and 𝒫¯dΘ\bar{\mathcal{P}}_{d}^{\Theta} carry the same (co)homological information. Let us describe it in pure combinatorial terms.

Let us consider the following domain in d+1\mathbb{R}^{d+1}:

d\displaystyle\mathcal{E}_{d} =𝖽𝖾𝖿\displaystyle=_{\mathsf{def}} {(u,P)×𝒫d|P(u)0} and its boundary\displaystyle\big{\{}(u,P)\in\mathbb{R}\times\mathcal{P}_{d}|\;P(u)\leq 0\big{\}}\text{\; and its boundary}
(2.2) d\displaystyle\partial\mathcal{E}_{d} =𝖽𝖾𝖿\displaystyle=_{\mathsf{def}} {(u,P)×𝒫d|P(u)=0}.\displaystyle\big{\{}(u,P)\in\mathbb{R}\times\mathcal{P}_{d}|\;P(u)=0\big{\}}.

The pair (d,d)(\mathcal{E}_{d},\partial\mathcal{E}_{d}) is of a fundamental importance for us.

For a subposet Θ𝛀\Theta\subset{\mathbf{\Omega}} and natural numbers d,kd,k, we introduce the following notations:

(2.3) Θ[d]=𝖽𝖾𝖿\displaystyle\Theta_{[d]}=_{\mathsf{def}} {ωΘ:|ω|=d},Θd]=𝖽𝖾𝖿{ωΘ:|ω|d},\displaystyle\{\omega\in\Theta:\;|\omega|=d\},\quad\Theta_{\langle d]}=_{\mathsf{def}}\{\omega\in\Theta:\;|\omega|\leq d\},
Θ||=k=𝖽𝖾𝖿\displaystyle\Theta_{|\sim|^{\prime}=k}=_{\mathsf{def}} {ωΘ:|ω|=k},Θ||k=𝖽𝖾𝖿{ωΘ:|ω|k}.\displaystyle\{\omega\in\Theta:\;|\omega|^{\prime}=k\},\quad\Theta_{|\sim|^{\prime}\geq k}=_{\mathsf{def}}\{\omega\in\Theta:\;|\omega|^{\prime}\geq k\}.

Assuming that Θ𝛀\Theta\subset{\mathbf{\Omega}} is a closed sub-poset, let

𝐜Θ=𝖽𝖾𝖿𝛀Θ and 𝐜Θd]=𝖽𝖾𝖿𝛀d]Θd].\mathbf{c}\Theta=_{\mathsf{def}}\mathbf{\Omega}\setminus\Theta\text{\; and \;}\mathbf{c}\Theta_{\langle d]}=_{\mathsf{def}}{\mathbf{\Omega}}_{\langle d]}\setminus\Theta_{\langle d]}.

We denote by [𝛀d]]\mathbb{Z}[{\mathbf{\Omega}}_{\langle d]}] the free \mathbb{Z}-module, generated by the elements of 𝛀d]{\mathbf{\Omega}}_{\langle d]}. Using the merge operators 𝖬k{\mathsf{M}}_{k} and the insert operators 𝖨k{\mathsf{I}}_{k} from [KSW2] on 𝛀{\mathbf{\Omega}}, we define homomorphisms 𝖬:[𝛀d]][𝛀d]]\partial_{{\mathsf{M}}}:\mathbb{Z}[{\mathbf{\Omega}}_{\langle d]}]\to\mathbb{Z}[{\mathbf{\Omega}}_{\langle d]}], 𝖨:[𝛀d]][𝛀d]]\partial_{{\mathsf{I}}}:\mathbb{Z}[{\mathbf{\Omega}}_{\langle d]}]\to\mathbb{Z}[{\mathbf{\Omega}}_{\langle d]}] by

𝖬(ω)=𝖽𝖾𝖿k=1sω1(1)k𝖬k(ω) and 𝖨(ω)=𝖽𝖾𝖿k=0sω(1)k𝖨k(ω),\partial_{{\mathsf{M}}}(\omega)=_{\mathsf{def}}-\sum_{k=1}^{s_{\omega}-1}(-1)^{k}{\mathsf{M}}_{k}(\omega)\,\text{ and }\,\partial_{{\mathsf{I}}}(\omega)=_{\mathsf{def}}\sum_{k=0}^{s_{\omega}}(-1)^{k}{\mathsf{I}}_{k}(\omega),

where sω=𝖽𝖾𝖿|ω||ω|s_{\omega}=_{\mathsf{def}}|\omega|-|\omega|^{\prime}. In fact, 𝖬\partial_{{\mathsf{M}}} and 𝖨\partial_{{\mathsf{I}}} are anti-commuting differentials [KSW2]. Therefore, the sum

(2.4) =𝖽𝖾𝖿𝖬+𝖨:[𝛀d]][𝛀d]]\displaystyle\partial=_{\mathsf{def}}\partial_{{\mathsf{M}}}+\partial_{{\mathsf{I}}}:\mathbb{Z}[{\mathbf{\Omega}}_{\langle d]}]\to\mathbb{Z}[{\mathbf{\Omega}}_{\langle d]}]

is a differential.

For a closed poset Θ\Theta, the restrictions of the operators ,𝖬\partial,\partial_{{\mathsf{M}}}, and 𝖨\partial_{{\mathsf{I}}} to the free \mathbb{Z}-module [Θ]\mathbb{Z}[\Theta] are well-defined. Thus, for any closed subposet Θd]𝛀d]\Theta_{\langle d]}\subset\mathbf{\Omega}_{\langle d]}, we may consider the differential complex :[Θd]][Θd]]\partial:\mathbb{Z}[\Theta_{\langle d]}]\to\mathbb{Z}[\Theta_{\langle d]}], whose (dj)(d-j)-grading is defined by the module [Θd]Θ||=j]\mathbb{Z}[\Theta_{\langle d]}\cap\Theta_{|\sim|^{\prime}=j}]. We denote by :[Θd]][Θd]]\partial^{\ast}:\mathbb{Z}[\Theta_{\langle d]}]^{\ast}\to\mathbb{Z}[\Theta_{\langle d]}]^{\ast} its dual operator, where [Θd]]=𝖽𝖾𝖿𝖧𝗈𝗆([Θd]],)\mathbb{Z}[\Theta_{\langle d]}]^{\ast}=_{\mathsf{def}}\mathsf{Hom}(\mathbb{Z}[\Theta_{\langle d]}],\mathbb{Z}).

Then we consider the quotient set 𝚯d]#:=𝛀d]/Θd]\mathbf{\Theta}^{\#}_{\langle d]}:=\mathbf{\Omega}_{\langle d]}\big{/}\Theta_{\langle d]}. For the closed subposet Θd]\Theta_{\langle d]}, the partial order in 𝛀d]\mathbf{\Omega}_{\langle d]} induces a partial order in the quotient 𝚯d]#\mathbf{\Theta}^{\#}_{\langle d]}.

Finally, we introduce a new differential complex ([𝚯d]#],#)(\mathbb{Z}[\mathbf{\Theta}^{\#}_{\langle d]}],\partial^{\#}) by including it in the short exact sequence of differential complexes:

(2.5) 0([Θd]],)([𝛀d]],)([𝚯d]#],#)0.\displaystyle 0\to(\mathbb{Z}[\Theta_{\langle d]}],\partial)\to(\mathbb{Z}[\mathbf{\Omega}_{\langle d]}],\partial)\to(\mathbb{Z}[\mathbf{\Theta}^{\#}_{\langle d]}],\partial^{\#})\to 0.

We will rely on the following result from [KSW2], which reduces the computation of the reduced cohomology H~(𝒫d𝐜Θ;)\tilde{H}^{\ast}(\mathcal{P}_{d}^{\mathbf{c}\Theta};\mathbb{Z}) to Algebra and Combinatorics.

Theorem 2.1.

([KSW2]) Let Θ𝛀d]\Theta\subset{\mathbf{\Omega}}_{\langle d]} be a closed subposet. Then, for any j[0,d]j\in[0,d], we get group isomorphisms

(2.6) H~j(𝒫d𝐜Θ;)𝒟Hdj(𝒫¯d,𝒫¯dΘ;)Hdj(#:[𝚯#][𝚯#]),\displaystyle\tilde{H}^{j}(\mathcal{P}_{d}^{\mathbf{c}\Theta};\mathbb{Z})\;\stackrel{{\scriptstyle\mathcal{D}}}{{\approx}}\;H_{d-j}(\bar{\mathcal{P}}_{d},\bar{\mathcal{P}}_{d}^{\Theta};\mathbb{Z})\;\approx H_{d-j}\big{(}\partial^{\#}:\mathbb{Z}[\mathbf{\Theta}^{\#}]\to\mathbb{Z}[\mathbf{\Theta}^{\#}]\big{)},
H~j(𝒫d𝐜Θ;)𝒟Hdj(𝒫¯d,𝒫¯dΘ;)Hdj((#):([𝚯#])([𝚯#])),\displaystyle\quad\tilde{H}_{j}(\mathcal{P}_{d}^{\mathbf{c}\Theta};\mathbb{Z})\;\stackrel{{\scriptstyle\mathcal{D}}}{{\approx}}\;H^{d-j}(\bar{\mathcal{P}}_{d},\bar{\mathcal{P}}_{d}^{\Theta};\mathbb{Z})\;\approx H_{d-j}\big{(}(\partial^{\#})^{\ast}:(\mathbb{Z}[\mathbf{\Theta}^{\#}])^{\ast}\to(\mathbb{Z}[\mathbf{\Theta}^{\#}])^{\ast}\big{)},

where 𝒟\mathcal{D} is the Poincaré duality isomorphism. \diamondsuit

Consider the embedding ϵd,d+2:𝒫d𝒫d+2\epsilon_{d,d+2}:\mathcal{P}_{d}\to\mathcal{P}_{d+2}, defined by the formula

(2.7) ϵd,d+2(P)(u)=𝖽𝖾𝖿(u2+1)P(u).\displaystyle\epsilon_{d,d+2}(P)(u)=_{\mathsf{def}}(u^{2}+1)\cdot P(u).

It preserves the 𝛀{\mathbf{\Omega}}-stratifications of the two spaces by the combinatorial types ω\omega of real divisors. The embedding ϵd,d+2\epsilon_{d,d+2} makes it possible to talk about stabilization of the homology/cohomology of the spaces 𝒫¯dΘ\bar{\mathcal{P}}_{d}^{\Theta} and 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta}, as dd\to\infty.

For a closed poset Θ𝛀\Theta\subset{\mathbf{\Omega}}, consider the closed finite poset Θd]=𝛀d]Θ\Theta_{\langle d]}={\mathbf{\Omega}}_{\langle d]}\cap\Theta. It is generated by some maximal elements ω(1),,ω()\omega_{\star}^{(1)},\dots,\omega_{\star}^{(\ell)}, where =(d)\ell=\ell(d). We introduce two useful quantities:

(2.8) ηΘ(d)\displaystyle\eta_{\Theta}(d) =𝖽𝖾𝖿\displaystyle=_{\mathsf{def}} maxi[1,(d)]{(|ω(i)|2|ω(i)|)},\displaystyle\max_{i\,\in\,[1,\,\ell(d)]}\big{\{}\big{(}|\omega_{\star}^{(i)}|-2|\omega_{\star}^{(i)}|^{\prime}\big{)}\big{\}},
(2.9) ψΘ(d)\displaystyle\psi_{\Theta}(d) =𝖽𝖾𝖿\displaystyle=_{\mathsf{def}} 12(d+ηΘ(d)).\displaystyle\frac{1}{2}\big{(}d+\eta_{\Theta}(d)\big{)}.

Note that

ψΘ(d)=12mini[1,(d)]{(d|ω(i)|)+(|ω(i)||ω(i)|)}<d,\psi_{\Theta}(d)=\frac{1}{2}\;\min_{i\,\in\,[1,\ell(d)]}\big{\{}(d-|\omega_{\star}^{(i)}|^{\prime})+(|\omega_{\star}^{(i)}|-|\omega_{\star}^{(i)}|^{\prime})\big{\}}<d,

where both summands, the “codimension” (d|ω(i)|)(d-|\omega_{\star}^{(i)}|^{\prime}) and the “support” (|ω(i)||ω(i)|)(|\omega_{\star}^{(i)}|-|\omega_{\star}^{(i)}|^{\prime}), are positive and each does not exceed dd. At the same time, ηΘ(d)\eta_{\Theta}(d) may be negative.

In the stabilization results about quasitopies, the quantity

(2.10) ξΘ(d+2)=𝖽𝖾𝖿d+2ψΘ(d+2).\displaystyle\xi_{\Theta}(d+2)=_{\mathsf{def}}d+2-\psi_{\Theta}(d+2).

plays the key role.

Now we are in position to state the main stabilization result from [KSW2]:

Theorem 2.2.

(short stabilization: {𝐝𝐝+𝟐}\mathbf{\{d\Rightarrow d+2\}}) Let Θ\Theta be a closed subposet of 𝛀\mathbf{\Omega}. Let the embedding ϵd,d+2:𝒫d𝒫d+2\epsilon_{d,d+2}:\,\mathcal{P}_{d}\subset\mathcal{P}_{d+2} be as in (2.7).

  • Then, for all jψΘ(d+2)1j\geq\psi_{\Theta}(d+2)-1, we get a homological isomorphism

    (ϵd,d+2):Hj(𝒫¯dΘ;)Hj+2(𝒫¯d+2Θ;),(\epsilon_{d,d+2})_{\ast}:H_{j}(\bar{\mathcal{P}}_{d}^{\Theta};\mathbb{Z})\approx H_{j+2}(\bar{\mathcal{P}}_{d+2}^{\Theta};\mathbb{Z}),
  • and, for all jd+2ψΘ(d+2)j\leq d+2-\psi_{\Theta}(d+2), a homological isomorphism

    (ϵd,d+2):Hj(𝒫d𝐜Θ;)Hj(𝒫d+2𝐜Θ;).(\epsilon_{d,d+2})_{\ast}:H_{j}(\mathcal{P}_{d}^{\mathbf{c}\Theta};\mathbb{Z})\approx H_{j}(\mathcal{P}_{d+2}^{\mathbf{c}\Theta};\mathbb{Z}).\quad\quad\diamondsuit
Definition 2.1.

A closed poset Θ𝛀\Theta\subseteq\mathbf{\Omega} is called profinite if, for all integers q0q\geq 0, there exist only finitely many elements ωΘ\omega\in\Theta such that |ω|q|\omega|^{\prime}\leq q. \diamondsuit

Corollary 2.1.

(long stabilization: {𝐝}\mathbf{\{d\Rightarrow\infty\}}) For any closed profinite poset Θ𝛀\Theta\subset\mathbf{\Omega}, and for each jj, the homomorphism

(ϵd,d):Hj(𝒫d𝐜Θ;)Hj(𝒫d𝐜Θ;)(\epsilon_{d,d^{\prime}})_{\ast}:\;H_{j}(\mathcal{P}_{d}^{\mathbf{c}\Theta};\mathbb{Z})\approx H_{j}(\mathcal{P}_{d^{\prime}}^{\mathbf{c}\Theta};\mathbb{Z})

is an isomorphism for all sufficiently big ddd\leq d^{\prime}, ddmod2d\equiv d^{\prime}\mod 2. \diamondsuit

As a result, we may talk about the stable homology Hj(𝒫𝐜Θ;)H_{j}(\mathcal{P}_{\infty}^{\mathbf{c}\Theta};\mathbb{Z}), the direct limit limdHj(𝒫d𝐜Θ;)\lim_{d\to\infty}H_{j}(\mathcal{P}_{d}^{\mathbf{c}\Theta};\mathbb{Z}).

Let us describe a few special cases of stabilization from [KSW2]. For k1k\geq 1, q[0,d]q\in[0,d], and qdmod2q\equiv d\mod 2, let us consider the closed poset

(2.11) 𝛀||k(q)=𝖽𝖾𝖿{ω𝛀d] such that |ω|k and |ω|q}.\displaystyle{\mathbf{\Omega}}_{|\sim|^{{}^{\prime}}\geq k}^{(q)}=_{\mathsf{def}}\,\big{\{}\omega\in{\mathbf{\Omega}}_{\langle d]}\text{ such that }|\omega|^{\prime}\geq k\text{ and }|\omega|\geq q\big{\}}.

Note that, for Θ=𝛀||k(0)=𝖽𝖾𝖿𝛀||k\Theta={\mathbf{\Omega}}_{|\sim|^{{}^{\prime}}\geq k}^{(0)}=_{\mathsf{def}}{\mathbf{\Omega}}_{|\sim|^{\prime}\geq k}, the space 𝒫¯dΘ\bar{\mathcal{P}}_{d}^{\Theta} is the entire (dk)(d-k)-skeleton of 𝒫¯d\bar{\mathcal{P}}_{d}.

(2.12) Let A(d,k,q)=𝖽𝖾𝖿|χ(([𝛀||k(q)],))|=|χ(𝒫¯d𝛀||k(q))1|,\displaystyle\text{ Let \;}A(d,k,q)=_{\mathsf{def}}\;\big{|}\chi\big{(}(\mathbb{Z}\big{[}{\mathbf{\Omega}}_{|\sim|^{{}^{\prime}}\geq k}^{(q)}\big{]},\partial)\big{)}\big{|}=\big{|}\chi\big{(}\bar{\mathcal{P}}_{d}^{{\mathbf{\Omega}}_{|\sim|^{{}^{\prime}}\geq k}^{(q)}}\big{)}-1\big{|},

the absolute value of the Euler number of the differential complex ([𝛀||k(q)],)\big{(}\mathbb{Z}\big{[}{\mathbf{\Omega}}_{|\sim|^{{}^{\prime}}\geq k}^{(q)}\big{]},\partial\big{)}.

Proposition 2.1.

([KSW2]) Fix k[1,d)k\in[1,d) and q0q\geq 0 such that qdmod2q\equiv d\mod 2. Let Θ=𝛀||k(q)\Theta={\mathbf{\Omega}}_{|\sim|^{{}^{\prime}}\geq k}^{(q)}.

Then the one-point compactification 𝒫¯dΘ\bar{\mathcal{P}}^{\Theta}_{d} has the homotopy type of a bouquet of (dk)(d-k)-dimensional spheres. The number of spheres in the bouquet equals A(d,k,q)A(d,k,q). \diamondsuit

Proposition 2.2.

([KSW1]) Let Θ𝛀d],||2\Theta\subseteq{\mathbf{\Omega}}_{\langle d],\,|\sim|^{\prime}\geq 2} be a closed poset. For dd+2d^{\prime}\geq d+2 such that ddmod2d^{\prime}\equiv d\mod 2, let Θ^{d}\hat{\Theta}_{\{d^{\prime}\}} be the smallest closed poset in 𝛀d]{\mathbf{\Omega}}_{\langle{d^{\prime}}]} containing Θ\Theta.

Then for dd+2d^{\prime}\geq d+2, we have an isomorphism π1(𝒫d𝐜Θ^{d})π1(𝒫d+2𝐜Θ{d+2})\pi_{1}\big{(}\mathcal{P}_{d^{\prime}}^{\mathbf{c}\hat{\Theta}_{\{d^{\prime}\}}}\big{)}\cong\pi_{1}\big{(}\mathcal{P}_{d+2}^{\mathbf{c}\Theta_{\{d+2\}}}\big{)} of the fundamental groups. \diamondsuit

3. Submersions & embeddings of manifolds whose boundary has constrained tangency patterns to the product 11-foliations

This section forms a bridge between the results from [K9] and our main results from Section 4. The reader may choose to surf Section 3 or to proceed directly to Section 4. Since all the results of this section are instant derivations of the similar results from [K9], we provide just an outline of their validations. Although we will not use directly the results of Section 3 in Section 4, this section may induce the “right mindset” for the reader.

Let YY be a smooth compact nn-manifold. Having in mind applications to traversing vector flows, we move away from immersions and embeddings β:M×Y\beta:M\to\mathbb{R}\times Y of nn-manifolds MM, the topic of [K9], to submersions and regular embeddings α:X×Y\alpha:X\to\mathbb{R}\times Y of compact smooth (n+1)(n+1)-manifolds XX with boundary into the product ×Y\mathbb{R}\times Y. When dim(X)=dim(Y)+1\dim(X)=\dim(Y)+1, the submersions and immersions are the same notion. Moreover, the restriction of a submersion α\alpha to X\partial X is an immersion. Of cause, if α:X×Y\alpha:X\to\mathbb{R}\times Y is an embedding, so is α:=α|:X×Y\alpha^{\partial}:=\alpha|:\partial X\to\mathbb{R}\times Y.

Therefore many constructions and notions from [K9], with the help of the correspondence αα\alpha\leadsto\alpha^{\partial}, apply instantly to submersions α:X×Y\alpha:X\to\mathbb{R}\times Y such that dimX=dimY+1\dim X=\dim Y+1.

Remark 3.1.

From the viewpoint of this paper, the main difference between immersions β:M×Y\beta:M\to\mathbb{R}\times Y and submersions α:X×Y\alpha:X\to\mathbb{R}\times Y, where dimX=dimM+1\dim X=\dim M+1, is that not any β\beta is a boundary α\alpha^{\partial} of some α\alpha. For example, the figure \infty in the plane does not bound a submersion of a 22-manifold. See [Pa] for the comprehensive theory of possible extensions of a given immersion β\beta to a submersion α\alpha. \diamondsuit

Example 3.1.

The following simple construction provides models of submersions that animate our treatment. Let WW be a codimension zero compact submanifold of a given manifold VV. Consider a covering map π:W~W\pi:\tilde{W}\to W with a finite fiber. Let XW~X\subset\tilde{W} be a compact codimension zero submanifold. It is possible to isotop the imbedding XW~X\subset\tilde{W} so that π:XW\pi:\partial X\to W will be an immersion with all the multiple crossings of π(X)\pi(\partial X) being in general position. Of course, each crossing has the multiplicity that does not exceed the cardinality of the π\pi-fiber. Then π:XV\pi:X\to V is the model example of a submersion to keep in mind. \diamondsuit

Let us introduce the central to this paper notion of quasitopy for submersions α:X×Y\alpha:X\to\mathbb{R}\times Y, an analogue of Definition 3.7 from [K9]. We fix a natural number dd and consider a closed sub-poset Θ𝛀\Theta\subset\mathbf{\Omega} such that ()Θ(\emptyset)\notin\Theta (the, so called, Λ\Lambda-condition (3.10) from [K9]). For topological reasons, we will consider only the case d0mod2d\equiv 0\mod 2.

Let \mathcal{L} be the 11-foliation of ×Y\mathbb{R}\times Y by the fibers of the obvious projection ×YY\mathbb{R}\times Y\to Y, and \mathcal{L}^{\bullet} be the 11-foliation of ×Y×[0,1]\mathbb{R}\times Y\times[0,1] by the fibers of ×Y×[0,1]Y×[0,1]\mathbb{R}\times Y\times[0,1]\to Y\times[0,1].

Let XX be a compact smooth (n+1)(n+1)-dimensional manifold with boundary. Consider a smooth map α:X×Y\alpha:X\to\mathbb{R}\times Y such that:

  • α\alpha is a submersion,

  • for each yYy\in Y, the total multiplicity mβ(y)m_{\beta}(y) of α(X)\alpha(\partial X) with respect to the foliation \mathcal{L} (see [K9], formula (3.4)) is less than or equal to dd and mβ(y)dmod2m_{\beta}(y)\equiv d\mod 2,

  • for each yYy\in Y, the combinatorial tangency pattern ωα(y)𝛀d]\omega^{\alpha}(y)\in\mathbf{\Omega}_{\langle d]} of α(X)\alpha(\partial X) with respect to \mathcal{L} does not belong to Θ\Theta,

  • for each yYy\in\partial Y, ωα(y)=()\omega^{\alpha}(y)=(\emptyset).

Note that the normal bundle να\nu^{\alpha} to α(X)\alpha(\partial X) in ×Y\mathbb{R}\times Y is trivial.

Let X0,X1X_{0},X_{1} be two compact smooth (oriented) (n+1)(n+1)-dimensional manifolds with boundary. We consider a compact smooth (oriented) (n+2)(n+2)-manifold WW with conners X0X1\partial X_{0}\coprod\partial X_{1} such that W=(X0X1){X0X1}δW\partial W=(X_{0}\coprod X_{1})\bigcup_{\{\partial X_{0}\coprod\partial X_{1}\}}\delta W, where δW\delta W is a smooth (oriented) cobordism between X0\partial X_{0} and X1\partial X_{1}. Let Z=𝖽𝖾𝖿×YZ=_{\mathsf{def}}\mathbb{R}\times Y and Z=𝖽𝖾𝖿×YZ^{\partial}=_{\mathsf{def}}\mathbb{R}\times\partial Y.

Let A:W×ZA:W\to\mathbb{R}\times Z, where dim(W)=dim(Z)+1\dim(W)=\dim(Z)+1, be a submersion. In particular, A:δW×Z,A|X0×(Y×{0})A:\delta W\to\mathbb{R}\times Z,\;A|_{\partial X_{0}}\to\mathbb{R}\times(Y\times\{0\}), and A|X1×(Y×{1})A|_{\partial X_{1}}\to\mathbb{R}\times(Y\times\{1\}) are immersions.

The next two definitions lay down the foundation for notions of quasitopy of traversing vector fields, the main subject of Section 4 (see Fig. 2).

Definition 3.1.

Let us fix natural numbers ddd^{\prime}\geq d, dd0mod2d^{\prime}\equiv d\equiv 0\mod 2. Consider closed subposets ΘΘ𝛀\Theta^{\prime}\subset\Theta\subset\mathbf{\Omega} such that ()Θ(\emptyset)\notin\Theta.

We say that a two submersions α0:X0×Y\alpha_{0}:X_{0}\to\mathbb{R}\times Y and α1:X1×Y\alpha_{1}:X_{1}\to\mathbb{R}\times Y are (d,d;𝐜Θ,𝐜Θ)(d,d^{\prime};\mathbf{c}\Theta,\mathbf{c}\Theta^{\prime})-quasitopic, if there exists a compact smooth (n+2)(n+2)-manifold WW as above and a smooth submersion A:W×ZA:W\to\mathbb{R}\times Z so that:

  • A|X0=α0A|_{X_{0}}=\alpha_{0} and A|X1=α1A|_{X_{1}}=\alpha_{1};

  • for each zZz\in Z, the total multiplicity mA(z)m_{A}(z) of A(δW)A(\delta W) with respect to the fiber z\mathcal{L}^{\bullet}_{z} is such that mA(z)dm_{A}(z)\leq d^{\prime}, mA(z)dmod2m_{A}(z)\equiv d^{\prime}\mod 2, and the combinatorial tangency pattern ωA(z)\omega^{A}(z) of A(δW)A(\delta W) with respect to z\mathcal{L}^{\bullet}_{z} belongs to 𝐜Θ\mathbf{c}\Theta^{\prime};

  • for each zY×({0}{1})z\in Y\times(\{0\}\cup\{1\}), the total multiplicity mA(z)m_{A}(z) of A(δW)A(\delta W) with respect to the fiber z\mathcal{L}_{z} is such that mA(z)dm_{A}(z)\leq d, mA(z)dmod2m_{A}(z)\equiv d\mod 2, and the combinatorial tangency pattern ωA(z)\omega^{A}(z) of A(δW)A(\delta W) with respect to \mathcal{L} belongs to 𝐜Θ\mathbf{c}\Theta.

  • for each zZz\in Z^{\partial}, ωA(z)=()\omega^{A}(z)=(\emptyset).

We denote by 𝖰𝖳¯¯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)\underline{\overline{\mathsf{QT}}}^{\mathsf{\,sub/emb}}_{\,d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) the set of quasitopy classes of such submersions/ embeddings α:X×Y\alpha:X\to\mathbb{R}\times Y. \diamondsuit

It is easy to check that the quasitopy of submersions is an equivalence relation. Recall that in [K9], Definition 3.7, we have introduced a similar notion of quasitopy for immersions/embeddings β:(M,M)(×Y,×Y)\beta:(M,\partial M)\to(\mathbb{R}\times Y,\mathbb{R}\times\partial Y). There, we used the notation

𝖰𝖳d,d𝗂𝗆𝗆/𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)=𝖽𝖾𝖿𝒬𝒯d,d𝗂𝗆𝗆/𝖾𝗆𝖻(Y,Y;𝐜Θ,();𝐜Θ)\mathsf{QT}^{\mathsf{imm/emb}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})=_{\mathsf{def}}\mathcal{QT}^{\mathsf{imm/emb}}_{d,d^{\prime}}(Y,\partial Y;\hfill\break\mathbf{c}\Theta,(\emptyset);\mathbf{c}\Theta^{\prime})

for the set of equivalence classes of immersions/embeddings β\beta under the quasitopy relation.

As for immersions β:M×Y\beta:M\to\mathbb{R}\times Y, for any choice of connected components κ1π0(Y1),κ2π0(Y2)\kappa_{1}\in\pi_{0}(\partial Y_{1}),\kappa_{2}\in\pi_{0}(\partial Y_{2}), the connected sum operation (see [K9], formula (3.15))

:𝖰𝖳¯¯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(Y1;𝐜Θ;𝐜Θ)×𝖰𝖳¯¯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(Y2;𝐜Θ;𝐜Θ)\displaystyle\uplus:\;\underline{\overline{\mathsf{QT}}}^{\mathsf{\,sub/emb}}_{\,d,d^{\prime}}(Y_{1};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\times\underline{\overline{\mathsf{QT}}}^{\mathsf{\,sub/emb}}_{\,d,d^{\prime}}(Y_{2};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\to
(3.2) 𝖰𝖳¯¯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(Y1#Y2;𝐜Θ;𝐜Θ),\displaystyle\to\underline{\overline{\mathsf{QT}}}^{\mathsf{\,sub/emb}}_{\,d,d^{\prime}}(Y_{1}\#_{\partial}Y_{2};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}),

is well-defined for the quasitopies of submersions.

It converts the set 𝖰𝖳¯¯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(Dn;𝐜Θ;𝐜Θ)\underline{\overline{\mathsf{QT}}}^{\mathsf{\,sub/emb}}_{\,d,d^{\prime}}(D^{n};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) into a group 𝖧¯¯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(n;𝐜Θ;𝐜Θ)\underline{\overline{\mathsf{H}}}^{\mathsf{\,sub/emb}}_{\,d,d^{\prime}}(n;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) (abelian for n>1n>1). This group acts, via the connected sum operation \uplus, on the set 𝖰𝖳¯¯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)\underline{\overline{\mathsf{QT}}}^{\mathsf{\,sub/emb}}_{\,d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}), provided that a connected component of Y\partial Y is chosen. To get a better insight, compare this claim with Proposition 3.2 from [K9].

For two pairs X1A1X_{1}\supset A_{1} and X2A2X_{2}\supset A_{2} of topological spaces, we denote by [(X1,A1),(X2,A2)][(X_{1},A_{1}),\hfill\break(X_{2},A_{2})] the set of homotopy classes of continuous maps g:X1X2g:X_{1}\to X_{2}, where g(A1)A2g(A_{1})\subset A_{2}.

Definition 3.2.

Given three pairs of spaces X1A1X_{1}\supset A_{1}, X2A2X_{2}\supset A_{2}, X3A3X_{3}\supset A_{3}, and a fixed continuous map ϵ:(X2,A2)(X3,A3)\epsilon:(X_{2},A_{2})\to(X_{3},A_{3}), we denote by

(3.3) [[(X1,A1),ϵ:(X2,A2)(X3,A3)]]\displaystyle[[(X_{1},A_{1}),\,\epsilon:(X_{2},A_{2})\to(X_{3},A_{3})]]

the set of homotopy classes [g][g] of continuous maps g:(X1,A1)(X2,A2)g:(X_{1},A_{1})\to(X_{2},A_{2}), modulo the following equivalence relation: by definition, [g0][g1][g_{0}]\sim[g_{1}], where g0:(X1,A1)(X2,A2)g_{0}:(X_{1},A_{1})\to(X_{2},A_{2}) and g1:(X1,A1)(X2,A2)g_{1}:(X_{1},A_{1})\to(X_{2},A_{2}) are continuous maps, if the compositions ϵg0\epsilon\circ g_{0} and ϵg1\epsilon\circ g_{1} are homotopic as maps from (X1,A1)(X_{1},A_{1}) to (X3,A3)(X_{3},A_{3}). \diamondsuit

Following the proof of Proposition 3.3 and Theorem 3.2 from [K9], we get Theorem 3.1, their analogue for submersions. It is crucial that here d0mod2d\equiv 0\mod 2, which implies that any regular embedding β:M×Y\beta:M\to\mathbb{R}\times Y, such that all the multiplicities {mβ(y)}yY\{m_{\beta}(y)\}_{y\in Y} are even, bounds a regular embedding α:X×Y\alpha:X\to\mathbb{R}\times Y, where X=M\partial X=M and α|X=β\alpha|_{\partial X}=\beta.

For an nn-dimensional YY, the next lemma reduces the computation of quasitopies 𝖰𝖳¯¯d,d𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)\underline{\overline{\mathsf{QT}}}^{\mathsf{\,emb}}_{\,d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}), based on regular embeddings α:X×Y\alpha:X\to\mathbb{R}\times Y of (n+1)(n+1)-dimensional manifolds XX, to the computation of quasitopies 𝖰𝖳d,d𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)\mathsf{QT}^{\mathsf{emb}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}), based on the regular embeddings β:M×Y\beta:M\to\mathbb{R}\times Y of nn-dimensional manifolds MM.

Lemma 3.1.

For closed posets ΘΘ𝛀\Theta^{\prime}\subset\Theta\subset\mathbf{\Omega} such that ()Θ(\emptyset)\notin\Theta and ddd\leq d^{\prime}, dd0mod2d\equiv d^{\prime}\equiv 0\mod 2, the map

Δ:𝖰𝖳¯¯d,d𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)𝖰𝖳d,d𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)\Delta:\,\underline{\overline{\mathsf{QT}}}^{\mathsf{\,emb}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\to\mathsf{QT}^{\mathsf{emb}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})

that takes an embedding α:X×Y\alpha:X\to\mathbb{R}\times Y to the embedding α:X×Y\alpha^{\partial}:\partial X\to\mathbb{R}\times Y is a bijection.

As a result, the obvious map 𝒜:𝖰𝖳¯¯d,d𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)𝖰𝖳¯¯d,d𝗂𝗆𝗆(Y;𝐜Θ;𝐜Θ)\mathcal{A}^{\bullet}:\,\underline{\overline{\mathsf{QT}}}^{\mathsf{\,emb}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\to\underline{\overline{\mathsf{QT}}}^{\mathsf{\,imm}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) is injective; in other words, if two embedding are quasitopic via a submersion, they are quasitopic via an embedding.

Proof.

The main step is contained in the proof of Lemma 3.6 from [K9]. Let us describe its flavor. Since, for d0mod2d\equiv 0\mod 2 and a closed nn-manifold MM, any (d)(\partial\mathcal{E}_{d})-regular (see Definition 3.4 in [K9]) embedding β:M×int(Y)\beta:M\subset\mathbb{R}\times\text{int}(Y) bounds a (orientable when YY is orientable) (n+1)(n+1)-manifold α:Xβ×Y\alpha:X_{\beta}\subset\mathbb{R}\times Y, the map Δ\Delta is onto. Evidently, the combinatorial tangency types to \mathcal{L} are determined by β(M)\beta(M). Thus, every embedding α𝖰𝖳¯¯d,d𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)\alpha\in\underline{\overline{\mathsf{QT}}}^{\mathsf{\,emb}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) produces an element α\alpha^{\partial} in 𝖰𝖳d,d𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)\mathsf{QT}^{\mathsf{\,emb}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}). By the same token, if a (d)(\partial\mathcal{E}_{d})-regular embedding β:M×int(Y)\beta:M\subset\mathbb{R}\times\text{int}(Y) bounds a (d)(\partial\mathcal{E}_{d^{\prime}})-regular embedding B:N×ZB:N\subset\mathbb{R}\times Z (where Z=𝖽𝖾𝖿Y×[0,1]Z=_{\mathsf{def}}Y\times[0,1]), whose tangency to \mathcal{L}^{\bullet} patterns belong to 𝐜Θ\mathbf{c}\Theta^{\prime}, then XβMNX_{\beta}\cup_{\partial M}N bounds (an orientable when YY is orientable) (n+2)(n+2)-manifold W×ZW\subset\mathbb{R}\times Z, provided d0mod2d^{\prime}\equiv 0\mod 2. Here N×ZN\subset\mathbb{R}\times Z is a compact (n+1)(n+1)-manifold such that N=M\partial N=M and B|M=βB|_{M}=\beta. Thus the map Δ\Delta is injective. By the previous argument, it is bijective.

By [K9], Proposition 3.5, the map 𝒜:𝖰𝖳d,d𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)𝖰𝖳d,d𝗂𝗆𝗆(Y;𝐜Θ;𝐜Θ)\mathcal{A}:\mathsf{QT}^{\mathsf{\,emb}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\to\mathsf{QT}^{\mathsf{\,imm}}_{d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) (see formula (3.32) in [K9]) is injective. By the argument above, Δ\Delta is injective. Chasing the obvious square diagram, formed by the sources and targets of the maps 𝒜,𝒜\mathcal{A},\mathcal{A}^{\bullet}, we conclude that 𝒜\mathcal{A}^{\bullet} is injective as well. ∎

Combining Lemma 3.1 with Theorem 3.2 from [K9], we get the following results.

Theorem 3.1.

We fix even natural numbers ddd^{\prime}\geq d, ddmod2d^{\prime}\equiv d\mod 2, and closed subposets ΘΘ𝛀d]\Theta^{\prime}\subset\Theta\subset\mathbf{\Omega}_{\langle d]} such that ()Θ(\emptyset)\notin\Theta. Let YY be a smooth compact nn-manifold.

Then any submersion α:X×Y\alpha:X\to\mathbb{R}\times Y as in (LABEL:multiplicity_condition) gives rise111not in a canonical fashion to a map Ψ(α):(Y,Y)(𝒫d𝐜Θ,𝒫d())\Psi(\alpha):(Y,\partial Y)\to(\mathcal{P}_{d}^{\mathbf{c}\Theta},\mathcal{P}_{d}^{(\emptyset)}). Moreover, (d,d;𝐜Θ,𝐜Θ)(d,d^{\prime};\mathbf{c}\Theta,\mathbf{c}\Theta^{\prime})-quasitopic submersions/embeddings α0:X0×Y\alpha_{0}:X_{0}\to\mathbb{R}\times Y and α1:X1×Y\alpha_{1}:X_{1}\to\mathbb{R}\times Y produce homotopic maps Ψ(α0)\Psi(\alpha_{0}) and Ψ(α1)\Psi(\alpha_{1}).

In this way, we get a map

Ψd,d𝗌𝗎𝖻/𝖾𝗆𝖻:𝖰𝖳¯¯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(Y;𝐜Θ;𝐜Θ)[[(Y,Y),ϵd,d:(𝒫d𝐜Θ,𝒫d())(𝒫d𝐜Θ,𝒫d())]],\Psi_{d,d^{\prime}}^{\mathsf{\,sub/emb}}:\;\underline{\overline{\mathsf{QT}}}^{\mathsf{\,sub/emb}}_{\,d,d^{\prime}}(Y;\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\to\big{[}\big{[}(Y,\partial Y),\;\epsilon_{d,d^{\prime}}:(\mathcal{P}_{d}^{\mathbf{c}\Theta},\mathcal{P}_{d}^{(\emptyset)})\to(\mathcal{P}_{d^{\prime}}^{\mathbf{c}\Theta^{\prime}},\mathcal{P}_{d^{\prime}}^{(\emptyset)})\big{]}\big{]},

Conversely, the homotopy class of any continuous map G:(Y,Y)(𝒫d𝐜Θ,𝒫d())G:(Y,\partial Y)\to(\mathcal{P}_{d}^{\mathbf{c}\Theta},\mathcal{P}_{d}^{(\emptyset)}) is realized by a smooth regular embedding α:X×Y\alpha:X\hookrightarrow\mathbb{R}\times Y which satisfies (LABEL:multiplicity_condition); that is, G=Ψd,d𝖾𝗆𝖻(α)G=\Psi_{d,d^{\prime}}^{\mathsf{\,emb}}(\alpha).

Moreover, Ψd,d𝖾𝗆𝖻\Psi_{d,d^{\prime}}^{\mathsf{\,emb}} is a bijection, and Ψd,d𝗌𝗎𝖻\Psi_{d,d^{\prime}}^{\mathsf{\,sub}} is a surjection, admitting a right inverse. \diamondsuit

4. Convex envelops of traversing flows, their quasitopies, & characteristic classes

4.1. Traversing, generic, and convex vector fields. Morse stratifications. Spaces of convex traversing vector fields

Definition 4.1.

A vector field v0v\neq 0 on a compact smooth manifold XX is called traversing, if each vv-trajectory is homeomorphic either to a closed interval, or to a point. \diamondsuit

Let vv be a traversing and boundary generic (see [K1], [K2], and Definition 4.4 below) vector field on a compact smooth (n+1)(n+1)-manifold XX with boundary. As we will see soon, every trajectory γ\gamma of such a vector field vv generates its tangency divisor DγD_{\gamma}, an ordered sequence of points in γ\gamma, together with their multiplicities (natural numbers).

We try to “go around” the fundamental discontinuity of the map xγxDγxx\to\gamma_{x}\to D_{\gamma_{x}}, where γx\gamma_{x} stands for the vv-trajectory through xXx\in X. This requires “to envelop” the pair (X,v)(X,v) in a convex envelop/pseudo-envelop (X^,v^)(\hat{X},\hat{v}) (see Definition 4.8). The convex pseudo-envelops, when available, will greatly simplify our analysis of traversing flows. In the spirit of Section 3, we will apply our results about immersions and submersions (against the background of product 11-foliations) from [K9] to the new environment of convex envelops of traversing flows.

Following [Mo], for a generic vector field vv on a smooth compact (n+1)(n+1)-dimensional manifold XX, such that v0v\neq 0 along X\partial X, let us describe an important Morse stratification {j±X(v)}j[1,dimX]\{\partial_{j}^{\pm}X(v)\}_{j\in[1,\dim X]} of the boundary X\partial X. The stratum jX=𝖽𝖾𝖿jX(v)\partial_{j}X=_{\mathsf{def}}\partial_{j}X(v) has the following description (see [K1]) in terms of an auxiliary function z:X^z:\hat{X}\to\mathbb{R} that satisfies the three properties:

  • 0 is a regular value of zz,

  • z1(0)=Xz^{-1}(0)=\partial X, and

  • z1((,0])=Xz^{-1}((-\infty,0])=X.

In terms of zz, the locus jX=𝖽𝖾𝖿jX(v)\partial_{j}X=_{\mathsf{def}}\partial_{j}X(v) is defined by the equations:

{z=0,vz=0,,v(j1)z=0},\big{\{}z=0,\;\mathcal{L}_{v}z=0,\;\ldots,\;\mathcal{L}_{v}^{(j-1)}z=0\big{\}},

where v(k)\mathcal{L}_{v}^{(k)} stands for the kk-th iteration of the Lie derivative operator v\mathcal{L}_{v} in the direction of vv (see [K2]). The pure stratum jXjX\partial_{j}X^{\circ}\subset\partial_{j}X is defined by the additional constraint v(j)z0\mathcal{L}_{v}^{(j)}z\neq 0. The locus jX\partial_{j}X is the union of two loci: (1) j+X\partial_{j}^{+}X, defined by the constraint v(j)z0\mathcal{L}_{v}^{(j)}z\geq 0, and (2) jX\partial_{j}^{-}X, defined by the constraint v(j)z0\mathcal{L}_{v}^{(j)}z\leq 0. The two loci, j+X\partial_{j}^{+}X and jX\partial_{j}^{-}X, share a common boundary j+1X\partial_{j+1}X.

For a generic vv, all the strata jX\partial_{j}X are smooth (n+1j)(n+1-j)-manifolds. The requirement of vv being generic with respect to X\partial X may be expressed as the property of the jj-form

(4.2) dzd(vz)d(v(j1)z)\displaystyle dz\wedge d(\mathcal{L}_{v}z)\wedge\;\ldots\;\wedge\;d(\mathcal{L}_{v}^{(j-1)}z)

being a nonzero section of the bundle jTX\bigwedge^{j}T_{\ast}X along the locus jX{\partial_{j}X}   for all j[1,n+1]j\in[1,n+1]. If vv on XX is generic to X\partial X, then each point xXx\in\partial X belongs to a unique minimal stratum jXX\partial_{j}X\subset\partial X with a maximal j=j(x)n+1j=j(x)\leq n+1. In the generic case, at each bXb\in\partial X, a flag

𝖥𝗅𝖺𝗀b(v)=𝖽𝖾𝖿{Tb(X)=FnFn1Fnj(b)+1}\mathsf{Flag}_{b}(v)={\mathsf{def}}\{T_{b}(\partial X)=F^{n}\supset F^{n-1}\supset\ldots\supset F^{n-j(b)+1}\}

is generated by the tangent spaces at bb to all the Morse strata {jX}jj(b)\{\partial_{j}X\}_{j\leq j(b)} that contain bb.

Let v^\hat{v} be a traversing vector field on a compact smooth (n+1)(n+1)-dimensional manifold X^\hat{X} with boundary. Consider a submersion α:X𝗂𝗇𝗍(X^)\alpha:X\to\mathsf{int}(\hat{X}), dimX=dimX^\dim X=\dim\hat{X}, such that the self-intersections of α(X)\alpha(\partial X) mutually transversal. Let v=α(v^)v=\alpha^{\dagger}(\hat{v}) be the transfer of v^\hat{v} to XX.

For general submersions α\alpha, which are not necessarily embeddings, the situation is more complex: not only one gets multiple self-intersections {Σk}k[2,n+1]\{\Sigma_{k}\}_{k\in[2,n+1]} of various branches of α(X)\alpha(\partial X), but such self-intersections may be tangent to the v^\hat{v}-flow in a variety of ways that produce similar Morse-type stratifications of the loci Σk\Sigma_{k}, k2k\geq 2, as well. Prior to Theorem 4.1, we will revisit this complication.

We associate several flags {𝖥𝗅𝖺𝗀b(v)}bα1(a)\{\mathsf{Flag}_{b}(v)\}_{b\in\alpha^{-1}(a)} with each point aα(X)a\in\alpha(\partial X). Let α\alpha^{\partial}_{\ast} denote the differential of the immersion α:XX^\alpha^{\partial}:\partial X\to\hat{X}.

Definition 4.2.

We say that several vector subspaces {ViW}i\{V_{i}\subset W\}_{i} of a given vector space WW are in general position, if the obvious map Wi(W/Vi)W\to\oplus_{i}(W/V_{i}) is onto. Note that this definition allows for any numbers of ViV_{i}’s to coincide with the ambient WW.

We say that the flags {(α)[𝖥𝗅𝖺𝗀b(v)]}bα1(a)\{(\alpha^{\partial}_{\ast})[\mathsf{Flag}_{b}(v)]\}_{b\in\alpha^{-1}(a)} are in general position in the ambient space TaX^T_{a}\hat{X}, if the α\alpha_{\ast}-images of the minimal strata {Fnj(b)+1(v)}bα1(a)\{F^{n-j(b)+1}(v)\}_{b\in\alpha^{-1}(a)} of the flags {𝖥𝗅𝖺𝗀b(v)}bα1(a)\{\mathsf{Flag}_{b}(v)\}_{b\in\alpha^{-1}(a)} are in general position in TaX^T_{a}\hat{X}. \diamondsuit

Example 4.1.

Consider the case n=2n=2, depicted in Fig.2.

If #(α)1(a)=1\#(\alpha^{\partial})^{-1}(a)=1, then each flag (α){F2F1F0}(\alpha^{\partial}_{\ast})\{F^{2}\supset F^{1}\supset F^{0}\}, or (α){F2F1}(\alpha^{\partial}_{\ast})\{F^{2}\supset F^{1}\}, or (α){F2}(\alpha^{\partial}_{\ast})\{F^{2}\} is in general position at aa.

If #(α)1(a)=2\#(\alpha^{\partial})^{-1}(a)=2, a pair of flags is in general position, if and only if, it is of the form (α){F2F1}(\alpha^{\partial}_{\ast})\{F^{2}\supset F^{1}\} and (α){G2}(\alpha^{\partial}_{\ast})\{G^{2}\}, or of the form (α){F2}(\alpha^{\partial}_{\ast})\{F^{2}\} and (α){G2}(\alpha^{\partial}_{\ast})\{G^{2}\}.

If #(α)1(a)=3\#(\alpha^{\partial})^{-1}(a)=3, a triple of flags is in general position only if the pair is of the form (α){F2}(\alpha^{\partial}_{\ast})\{F^{2}\}, (α){G2}(\alpha^{\partial}_{\ast})\{G^{2}\}, (α){H2}(\alpha^{\partial}_{\ast})\{H^{2}\}. The rest of combinations fail to be generic. \diamondsuit

Refer to caption


Figure 2. Six locally generic configurations of α(X)\alpha^{\partial}(\partial X) in 3D. The numbers 1,2,31,2,3 reflect the local multiplicity of the marked point aa on the trajectory γ^a\hat{\gamma}_{a} through it.
Definition 4.3.

Let X,X^X,\hat{X} be smooth compact (n+1)(n+1)-manifolds with boundary and v^\hat{v} a traversing vector field on X^\hat{X}. We assume that a submersion α:XX^\alpha:X\to\hat{X} has the following properties: for each point aα(X)a\in\alpha(\partial X), there exist a natural number k=k(a)n+1k=k(a)\leq n+1, an open neighborhood UaU_{a} of aa in X^\hat{X}, and smooth functions {z1,,zk:Ua}\{z_{1},\dots,z_{k}:U_{a}\to\mathbb{R}\} such that:

  1. (1)

    0 is a regular value for each ziz_{i},

  2. (2)

    in UaU_{a}, the locus α(X)\alpha(\partial X) is given by the equation {z1zk=0}\{z_{1}\cdot\ldots\cdot z_{k}=0\},

  3. (3)

    the differential kk-form dz1dzk|Ua0.dz_{1}\wedge\ldots\wedge dz_{k}\,|_{U_{a}}\neq 0.222Thus α\alpha^{\partial} is kk-normal in the sense of Definition 3.3 from [K9].

Let γ^a\hat{\gamma}_{a} denote the v^\hat{v}-trajectory through aa.

We say that a point aα(X)a\in\alpha(\partial X) has a multiplicity j=j(a)j=j(a) with respect to v^\hat{v}, if the jet 𝗃𝖾𝗍aj1(z1zk|γ^a)=0\mathsf{jet}^{j-1}_{a}\big{(}z_{1}\cdot\ldots\cdot z_{k}\big{|}_{\hat{\gamma}_{a}}\big{)}=0, but 𝗃𝖾𝗍aj(z1zk|γ^a)0\mathsf{jet}^{j}_{a}\big{(}z_{1}\cdot\ldots\cdot z_{k}\big{|}_{\hat{\gamma}_{a}}\big{)}\neq 0. \diamondsuit

Definition 4.4.

Let X,X^X,\hat{X} be smooth compact (n+1)(n+1)-manifolds with boundary and v^\hat{v} a traversing vector field on X^\hat{X}. Let UU be an open v^\hat{v}-flow adjusted neighborhood of aa in X^\hat{X}. For each point a𝗂𝗇𝗍(X^)a\in\mathsf{int}(\hat{X}), consider a smooth transversal section SS of the v^\hat{v}-flow at aa and the flow-generated local projections π:US\pi:U\to S.

We say that a submersion α:X𝗂𝗇𝗍(X^)\alpha:X\to\mathsf{int}(\hat{X}) is locally generic relative to v^\hat{v} if, for each point aα(X)a\in\alpha(\partial X),

  • the images of the flags333equivalently, of their minimal strata {𝖥𝗅𝖺𝗀b(v)}bα1(a)\{\mathsf{Flag}_{b}(v)\}_{b\in\alpha^{-1}(a)}, under the differentials (α)(\alpha^{\partial})_{\ast}, are in general position in Ta(X^)T_{a}(\hat{X}),

  • the images of the flags {𝖥𝗅𝖺𝗀b(v)}bα1(a)\{\mathsf{Flag}_{b}(v)\}_{b\in\alpha^{-1}(a)}, under the differentials (πα)(\pi\circ\alpha^{\partial})_{\ast}, are in general position in the tangent space TaST_{a}S. \diamondsuit

One may compare the next definition, which utilizes the notion of convexity, with Definition 4.6, introducing the more general notion of kk-convexity.

Definition 4.5.

Let X^\hat{X} be a compact connected smooth manifold with boundary, equipped with a vector field v^\hat{v}. We say that the pair (X^,v^)(\hat{X},\hat{v}) is convex if

  • v^\hat{v} admits a Lyapunov function f^:X^\hat{f}:\hat{X}\to\mathbb{R} (i.e., df^(v^)>0d\hat{f}(\hat{v})>0); equivalently, v^\hat{v} is traversing,

  • X^\partial\hat{X} is locally generic with respect to v^\hat{v},

  • 2+X^(v^)=\partial_{2}^{+}\hat{X}(\hat{v})=\emptyset (equivalently, v^\hat{v} is convex). \diamondsuit

Remark 4.1.

Any compact connected manifold X^\hat{X} with boundary has a traversing vector field [K1]. However, not any compact manifold with boundary admits a traversing convex vector field! For example, consider any surface X^\hat{X}, obtained from a closed oriented connected surface, different from the 22-sphere, by removing an open disk. Such an X^\hat{X} does not admit a traversing convex vector field [K5].

By [K1], Lemma 4.1, any traversing vector field v^\hat{v}, admits a Lyapunov function.

If YY is a closed manifold, then any vector field v^0\hat{v}\neq 0 that is tangent to the fibers of the obvious projection [0,1]×YY[0,1]\times Y\to Y is evidently convex with respect to [0,1]×Y\partial[0,1]\times Y. The obvious function f^:[0,1]×Y[0,1]\hat{f}:[0,1]\times Y\to[0,1] has the desired Lyapunov property df^(v^)>0d\hat{f}(\hat{v})>0. \diamondsuit

Example 4.2.

Any non-trapping Riemannian metric gg on a compact smooth manifold MM with a convex boundary M\partial M produces the geodesic vector field v^g\hat{v}^{g} on the unit spherical bundle SMMSM\to M, which is traversing and convex with respect to (SM)\partial(SM) [K6]. The very existence of such a metric gg puts severe restrictions on the topological nature of MM. \diamondsuit

Let 𝖼𝗈𝗇𝗏(X^)\mathsf{conv}(\hat{X}) denote the space of traversing vector fields v^\hat{v} on X^\hat{X} such that (X^,v^)(\hat{X},\hat{v}) is a convex pair, as in Definition 4.5. The space 𝖼𝗈𝗇𝗏(X^)\mathsf{conv}(\hat{X}) is considered in the CC^{\infty}-topology.

Lemma 4.1.

For any v^0,v^1𝖼𝗈𝗇𝗏(X^)\hat{v}_{0},\hat{v}_{1}\in\mathsf{conv}(\hat{X}) that belong to the same path-connected component of the space 𝖼𝗈𝗇𝗏(X^)\mathsf{conv}(\hat{X}), there exists a smooth diffeomorphism ϕ:X^X^\phi:\hat{X}\to\hat{X} such that ϕ\phi maps v^0\hat{v}_{0}-trajectories to v^1\hat{v}_{1}-trajectories, while preserving their orientations.

If, for a pair v^0,v^1𝖼𝗈𝗇𝗏(X^)\hat{v}_{0},\hat{v}_{1}\in\mathsf{conv}(\hat{X}), there exists a smooth isotopy {ψt:X^X^}t[0,1]\{\psi^{t}:\partial\hat{X}\to\partial\hat{X}\}_{t\in[0,1]} such that ψ1(1+X^(v^0))=1+X^(v^1)\psi^{1}(\partial_{1}^{+}\hat{X}(\hat{v}_{0}))=\partial_{1}^{+}\hat{X}(\hat{v}_{1}), then exists a smooth diffeomorphism ϕ:X^X^\phi:\hat{X}\to\hat{X}, an extension of ψ1\psi^{1}, that maps v^0\hat{v}_{0}-trajectories to v^1\hat{v}_{1}-trajectories, while preserving their orientations.

Proof.

If v^𝖼𝗈𝗇𝗏(X^)\hat{v}\in\mathsf{conv}(\hat{X}) is a convex boundary generic vector field, then by Theorem 6.6 from [K7], the stratification X^1+X^(v^)2X^(v^)\hat{X}\supset\partial_{1}^{+}\hat{X}(\hat{v})\supset\partial_{2}^{-}\hat{X}(\hat{v}) is stable, up to an isotopy of X^\hat{X}, under sufficiently small perturbations of v^\hat{v}. As a result, within a path connected component of v^\hat{v} in 𝖼𝗈𝗇𝗏(X^)\mathsf{conv}(\hat{X}), the smooth topological type of this Morse stratification is stable via an isotopy. In particular, 1+X^(v^0)\partial_{1}^{+}\hat{X}(\hat{v}_{0}) and 1+X^(v^1)\partial_{1}^{+}\hat{X}(\hat{v}_{1}) are isotopic in X^\hat{X}, provided the v^0\hat{v}_{0} and v^1\hat{v}_{1} are connected by a path in 𝖼𝗈𝗇𝗏(X^)\mathsf{conv}(\hat{X}).

Let us denote by {ψ^t:X^X^}t[0,1]\{\hat{\psi}^{t}:\hat{X}\to\hat{X}\}_{t\in[0,1]} the isotopy that transforms the v^0\hat{v}_{0}-induced Morse stratification of X^\partial\hat{X} to the v^1\hat{v}_{1}-induced Morse stratification of X^\partial\hat{X}. Let us compare the vector fields v~0=𝖽𝖾𝖿(ψ^1)(v^0)\tilde{v}_{0}=_{\mathsf{def}}(\hat{\psi}^{1})_{\ast}(\hat{v}_{0}) and v^1\hat{v}_{1}. Both vector fields point inside of X^\hat{X} exactly along 1+X^(v^1)\partial_{1}^{+}\hat{X}(\hat{v}_{1}). Since v^0,v^1\hat{v}_{0},\hat{v}_{1} are traversing, they admit some Lyapunov functions f0,f1:X^f_{0},f_{1}:\hat{X}\to\mathbb{R}. Put f~0=𝖽𝖾𝖿((ψ^1)1)(f0)\tilde{f}_{0}=_{\mathsf{def}}((\hat{\psi}^{1})^{-1})^{\ast}(f_{0}). It serves as Lyapunov’s function for v~0\tilde{v}_{0}. For x1+X^(v^1)x\in\partial_{1}^{+}\hat{X}(\hat{v}_{1}), we denote by γx{0}\gamma_{x}^{\{0\}} and γx{1}\gamma_{x}^{\{1\}} the v~0\tilde{v}_{0}- and v^1\hat{v}_{1}-trajectories through xx. Let var0(x)var_{0}(x) stands for the variation of the function f~0\tilde{f}_{0} along γx{0}\gamma_{x}^{\{0\}} and var1(x)var_{1}(x) for the variation of the function f^1\hat{f}_{1} along γx{1}\gamma_{x}^{\{1\}}.

For yγx{0}y\in\gamma_{x}^{\{0\}}, consider the unique point ϕ~(y)γx{1}\tilde{\phi}(y)\in\gamma_{x}^{\{1\}} such that f1(ϕ~(y))/var1(x)=f~0(y)/var0(x)f_{1}(\tilde{\phi}(y))/var_{1}(x)=\tilde{f}_{0}(y)/var_{0}(x). Now the diffeomorphism ϕ:=ϕ~ψ^1\phi:=\tilde{\phi}\circ\hat{\psi}^{1} takes v^0\hat{v}_{0}-trajectories to v^0\hat{v}_{0}-trajectories, while preserving their orientations. Hence, we have shown that the isotopy class of 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}) in X\partial X determines the smooth topological type of a convex pair (X^,(v^))(\hat{X},\mathcal{L}(\hat{v})), where (v^)\mathcal{L}(\hat{v}) denotes the oriented 11-foliation, determined by a convex traversing v^\hat{v}. ∎

Lemma 4.2.

Let X^\hat{X} is a connected compact smooth manifold with boundary. If #(π0(X^))3\#(\pi_{0}(\partial\hat{X}))\geq 3, then 𝖼𝗈𝗇𝗏(X^)=\mathsf{conv}(\hat{X})=\emptyset.

Proof.

For a convex v^\hat{v}, 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}) and 1X^(v^)\partial_{1}^{-}\hat{X}(\hat{v}), each is a deformation retract of X^\hat{X}. Therefore, each of these two loci must be connected. Thus each of the loci 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}), 1X^(v^)\partial_{1}^{-}\hat{X}(\hat{v}) must be contained in some connected component of X^\partial\hat{X}. When #(π0(X^))3\#(\pi_{0}(\partial\hat{X}))\geq 3, this argument forces at least one component of the boundary to be free from both loci. However, the union of the two loci is the entire boundary. This contradiction proves the claim. ∎

4.2. Homology spheres and convex flows

Let 𝐇𝐒n1\mathbf{HS}_{n-1} denote the monoid of smooth types of integral homology (n1)(n-1)-spheres, being considered up to connected sums with smooth homotopy (n1)(n-1)-spheres.

For n4n\neq 4, let 𝚯n\mathbf{\Theta}_{n} denote the group of hh-cobordism classes of smooth homotopy nn-spheres. The operations in 𝐇𝐒n1\mathbf{HS}_{n-1} and in 𝚯n\mathbf{\Theta}_{n} are the connected sums of spheres. We denote the order of the group 𝚯n\mathbf{\Theta}_{n} by |𝚯n||\mathbf{\Theta}_{n}|.

Proposition 4.1.

Any convex vector field v^\hat{v} on the standard ball Dn+1D^{n+1} defines a smooth involution τv^\tau_{\hat{v}} on Sn=Dn+1S^{n}=\partial D^{n+1}, whose fixed point set 2Dn+1(v^)\partial_{2}^{-}D^{n+1}(\hat{v}) is an integral homology (n1)(n-1)-sphere.

For n6n\geq 6, any element of the set 𝐇𝐒n1\mathbf{HS}_{n-1} arises as the locus 2X^(v^)\partial_{2}^{-}\hat{X}(\hat{v}) for a convex traversing vector field v^\hat{v} on a smooth compact contractible (n+1)(n+1)-dimensional manifold X^\hat{X}, whose boundary X^\partial\hat{X} is a smooth homotopy sphere. Moreover, X^\partial\hat{X} admits a smooth involution τv^\tau_{\hat{v}} such that (X^)τv^=2X^(v^)(\partial\hat{X})^{\tau_{\hat{v}}}=\partial_{2}^{-}\hat{X}(\hat{v}).

For n6n\geq 6, the multiple |𝚯n|[Σ]|\mathbf{\Theta}_{n}|\cdot[\Sigma] of any given element [Σ]𝐇𝐒n1[\Sigma]\in\mathbf{HS}_{n-1} arises as the locus 2Dn+1(v^)\partial_{2}^{-}D^{n+1}(\hat{v}) for a convex traversing vector field v^\hat{v} on the ball Dn+1D^{n+1}. The sphere Dn+1\partial D^{n+1} admits a smooth involution τv^\tau_{\hat{v}} such that (Dn+1)τv^=2Dn+1(v^)(\partial D^{n+1})^{\tau_{\hat{v}}}=\partial_{2}^{-}D^{n+1}(\hat{v}).

Proof.

Using a convex v^\hat{v}-flow, 1+Dn+1(v^)\partial_{1}^{+}D^{n+1}(\hat{v}) is a deformation retract of Dn+1D^{n+1} and thus a contractible manifold. By Poincaré duality, 2Dn+1(v^)\partial_{2}^{-}D^{n+1}(\hat{v}), the boundary of 1+Dn+1(v^)\partial_{1}^{+}D^{n+1}(\hat{v}), is a homology sphere.

By [Ke], Theorem 3, for n15n-1\geq 5, any smooth homology sphere Σn1\Sigma^{n-1}, after a connected sum Σn1#ΣHn1\Sigma^{n-1}\#\,\Sigma^{n-1}_{H} with a unique smooth homotopy (n1)(n-1)-sphere ΣHn1\Sigma^{n-1}_{H}, bounds a contractible smooth manifold WnW^{n}.

Consider a smooth metric gg on WnW^{n}. We denote by dg(x,Wn)d_{g}(x,\partial W^{n}) the smooth distance function to Wn\partial W^{n} on a collar UU of Wn\partial W^{n} in WnW^{n}. Let F:Wn+F:W^{n}\to\mathbb{R}_{+} be a function that is strictly positive and smooth in the interior of WnW^{n} and coincides with the function F~(x):=dg(x,Wn)\tilde{F}(x):=\sqrt{d_{g}(x,\partial W^{n})} in the collar UU.

Consider a smooth manifold X^n+1×Wn\hat{X}^{n+1}\subset\mathbb{R}\times W^{n}, given by the inequality {(t,x):|t|F(x)}\{(t,x):|t|\leq F(x)\}. It comes with the vector field v^\hat{v} that is tangent to the fibers of the obvious projection p:×WnWnp:\mathbb{R}\times W^{n}\to W^{n}. Since WnW^{n} is contractible, the boundary of X^n+1\hat{X}^{n+1}, the double of WnW^{n}, is a smooth homotopy nn-sphere, and X^n+1\hat{X}^{n+1} is a homotopy ball. The vector field v^\hat{v} on X^n+1\hat{X}^{n+1} is convex and defines an involution τv^:X^n+1X^n+1\tau_{\hat{v}}:\partial\hat{X}^{n+1}\to\partial\hat{X}^{n+1}, whose fixed point set (X^n+1)τv^=2X^n+1(v^)=W=Σn1#ΣHn1(\partial\hat{X}^{n+1})^{\tau_{\hat{v}}}=\partial_{2}^{-}\hat{X}^{n+1}(\hat{v})=\partial W=\Sigma^{n-1}\#\,\Sigma^{n-1}_{H}. This validates the second claim.

The connected sum of |𝚯n||\mathbf{\Theta}_{n}| copies of X^n+1\partial\hat{X}^{n+1} is a standard nn-sphere. Consider the boundary connected sum (Y^n+1,v~)(\hat{Y}^{n+1},\tilde{v}) of the |𝚯n||\mathbf{\Theta}_{n}| copies of the pair (X^n+1,v^)(\hat{X}^{n+1},\hat{v}). Here the 11-handles H=Dn×[0,1]H=D^{n}\times[0,1] are attached at pairs of points that belong to different pairs of Wn\partial W^{n}’s so that |𝚯n||\mathbf{\Theta}_{n}| copies of WnW^{n} are connected by the chain of 11-handles D+n1×[0,1]D^{n-1}_{+}\times[0,1], where D+n1DnD^{n-1}_{+}\subset\partial D^{n} is a hemisphere. The fields v~\tilde{v} in the different copies extend concavely across the 11-handles. Then Y^n+1\partial\hat{Y}^{n+1} is the standard sphere SnS^{n} which bounds a contractible manifold Y^n+1\hat{Y}^{n+1}. By the hh-cobordism theorem (see [Mi]), applied to Y^n+1Dn+1\hat{Y}^{n+1}\setminus D^{n+1}, we conclude that Y^n+1\hat{Y}^{n+1} is the standard ball. Thus we managed to build a convex vector field v#v^{\#} on Dn+1D^{n+1} whose locus 2Dn+1(v#)\partial_{2}^{-}D^{n+1}(v^{\#}) is a homology sphere Σ~n1\tilde{\Sigma}^{n-1}, the |𝚯n||\mathbf{\Theta}_{n}|-multiple of the given class Σn1𝐇𝐒n1\Sigma^{n-1}\in\mathbf{HS}_{n-1}. In particular, Dn+1\partial D^{n+1} admits a smooth involution τv#\tau_{v^{\#}} whose fixed point set is |𝚯n|Σn1|\mathbf{\Theta}_{n}|\cdot\Sigma^{n-1}. ∎

Example 4.3.

Consider a free action of the icosahedral group 𝖨120\mathsf{I}_{120} on S3S^{3}. Then 𝖨120\mathsf{I}_{120} acts freely on S7=𝗃𝗈𝗂𝗇(S3,S3)S^{7}=\mathsf{join}(S^{3},S^{3}). Hence the orbit space Σ7=𝖽𝖾𝖿S7/𝖨120\Sigma^{7}=_{\mathsf{def}}S^{7}/\mathsf{I}_{120} is a homology sphere. By Proposition 4.1, there is a convex traversing vector field v^\hat{v} on the ball D9D^{9}, such that its locus 2D9(v^)\partial_{2}^{-}D^{9}(\hat{v}) is a connected sum of 2828 copies of [Σ7]𝐇𝐒7[\Sigma^{7}]\in\mathbf{HS}_{7}, and its locus 1+D9(v^)\partial_{1}^{+}D^{9}(\hat{v}) is contractible. Note that π1(2D9(v^))\pi_{1}(\partial_{2}^{-}D^{9}(\hat{v})) is a free product of 28 copies of 𝖨120\mathsf{I}_{120}. \diamondsuit

Corollary 4.1.

For n>6n>6, π0(𝖼𝗈𝗇𝗏(Dn+1))\pi_{0}(\mathsf{conv}(D^{n+1})) admits a surjection onto a subgroup 𝐆n1\mathbf{G}_{n-1} of 𝐇𝐒n1\mathbf{HS}_{n-1} of index |𝚯n||\mathbf{\Theta}_{n}| at most.

In particular, here are a few “clean” surjections:

π0(𝖼𝗈𝗇𝗏(D7))𝐇𝐒5,π0(𝖼𝗈𝗇𝗏(D13))𝐇𝐒11,π0(𝖼𝗈𝗇𝗏(D62))𝐇𝐒60.\pi_{0}(\mathsf{conv}(D^{7}))\to\mathbf{HS}_{5},\;\pi_{0}(\mathsf{conv}(D^{13}))\to\mathbf{HS}_{11},\;\pi_{0}(\mathsf{conv}(D^{62}))\to\mathbf{HS}_{60}.
Proof.

For n15n-1\geq 5, let Σn1\Sigma^{n-1} be a given smooth homology sphere, and let ΣHn1\Sigma^{n-1}_{H} be the unique homotopy sphere such that Σn1#ΣHn1\Sigma^{n-1}\#\,\Sigma^{n-1}_{H} bounds a smooth contractible manifold [Ke]. By Proposition 4.1, any convex traversing vector field v^\hat{v} produces a homology sphere 2Dn+1(v^)\partial_{2}^{-}D^{n+1}(\hat{v}), and the |𝚯n||\mathbf{\Theta}_{n}|-multiple of any Σn1#ΣHn1\Sigma^{n-1}\#\,\Sigma^{n-1}_{H} is produced this way. On the other hand, deforming v^\hat{v} within the space of boundary generic vector fields does not change the smooth isotopy type of the pair 1+Dn+1(v^)2+Dn+1(v^)\partial_{1}^{+}D^{n+1}(\hat{v})\supset\partial_{2}^{+}D^{n+1}(\hat{v}) (see Lemma 4.1, or [K1], [K2]). In particular, the smooth topological type of the pair is preserved along a path in the space 𝖼𝗈𝗇𝗏(Dn+1)\mathsf{conv}(D^{n+1}). Therefore, π0(𝖼𝗈𝗇𝗏(Dn+1))\pi_{0}(\mathsf{conv}(D^{n+1})) admits a surjection onto a subgroup 𝐆n1\mathbf{G}_{n-1} of 𝐇𝐒n1\mathbf{HS}_{n-1} of index |𝚯n||\mathbf{\Theta}_{n}| at most.

The three examples of surjections in the corollary are based on the computations of |𝚯n||\mathbf{\Theta}_{n}| in [KeM], [WaX]; we just use some nn’s for which |𝚯n|=1|\mathbf{\Theta}_{n}|=1. ∎

4.3. Convex pseudo-envelops of traversing flows

The next key lemma incapsulates a given convex traversing flow (X^,v^)(\hat{X},\hat{v}) into the obvious traversing flow v~\tilde{v} in a box [0,1]×Y[0,1]\times Y for an appropriate choice of a compact manifold YY, dimY=dimX^\dim Y=\dim\partial\hat{X}. The construction that realizes the embedding (X^,v^)([0,1]×Y,v~)(\hat{X},\hat{v})\subset([0,1]\times Y,\tilde{v}) delivers a global “virtual section” {0}×Y\{0\}\times Y of v^\hat{v}. In turn, this section enables us to apply the key Theorem 3.1 from [K9] to any convex traversing flow (X^,v^)(\hat{X},\hat{v}). Therefore, we will be able to transfer many results from [K9] and from Section 3 to the environment of convex envelops and pseudo-envelops (see Definition 4.5) of traversing boundary generic vector fields.

Lemma 4.3.

Let dim(X^)=n+1\dim(\hat{X})=n+1. If a pair (X^,v^)(\hat{X},\hat{v}) is convex, then there exists a compact smooth nn-manifold YY such that:

  1. (1)

    X^[0,1]×Y\hat{X}\subset[0,1]\times Y,

  2. (2)

    v^\hat{v} is tangent to the fibers of the projection p:[0,1]×YYp:[0,1]\times Y\to Y,

  3. (3)

    the obvious function h:[0,1]×Y[0,1]h:[0,1]\times Y\to[0,1] has the property dh(v^)>0dh(\hat{v})>0,

  4. (4)

    with the help of pp, the loci 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}) and 1X^(v^)\partial_{1}^{-}\hat{X}(\hat{v}) each is homeomorphic to YY.

Proof.

Since vv is a traversing field, it admits a Lyapunov function f^:X^\hat{f}:\hat{X}\to\mathbb{R} so that df^(v)>0d\hat{f}(v)>0 in X^\hat{X} [K1]. We add a collar UU to X^\hat{X} along its boundary X^\partial\hat{X} and denote X^X^U\hat{X}\cup_{\partial\hat{X}}U by X~\tilde{X}. Then, we smoothly extend v^\hat{v} and f^\hat{f} in X~\tilde{X} and denote these extensions by v~\tilde{v} and f~\tilde{f}. We adjust UU so that v~0\tilde{v}\neq 0 and f~(v~)>0\tilde{f}(\tilde{v})>0 there.

Let F:X~F:\tilde{X}\to\mathbb{R} be a smooth function such that 0 is its regular value and F1(0)=X^F^{-1}(0)=\partial\hat{X}, F1((,0])=X^F^{-1}((-\infty,0])=\hat{X}. Let X~(F,ϵ)=𝖽𝖾𝖿{xX~|F(x)ϵ}\tilde{X}(F,\epsilon)=_{\mathsf{def}}\{x\in\tilde{X}|\;F(x)\leq\epsilon\}. For a sufficiently small ϵ>0\epsilon>0, X~(F,ϵ)\tilde{X}(F,\epsilon) is a smooth compact manifold, contained in X~\tilde{X}.

By definition, 2X^(v~):={xX^|F(x)=0, and (v~F)(x)=0}\partial_{2}\hat{X}(\tilde{v}):=\{x\in\partial\hat{X}|\;F(x)=0,\text{ and }(\mathcal{L}_{\tilde{v}}F)(x)=0\}. The convexity of v^\hat{v} in X^\hat{X} means that 2+X^(v^)=\partial_{2}^{+}\hat{X}(\hat{v})=\emptyset, thus 2X^(v^)=2X^(v^)\partial_{2}^{-}\hat{X}(\hat{v})=\partial_{2}\hat{X}(\hat{v}). By Morin’s Theorem [Mor] (see [K2] for details), in the vicinity of each point x2X^(v~)x\in\partial_{2}^{-}\hat{X}(\tilde{v}) there is a system of smooth coordinates (u,w,y1,,yn1)(u,w,y_{1},\ldots,y_{n-1}) in which F((u,w,y1,,yn1))=u2+wF((u,w,y_{1},\ldots,y_{n-1}))=u^{2}+w, so that X^\partial\hat{X} is given by the equation u2+w=0u^{2}+w=0, X^\hat{X} by the inequality u2+w0u^{2}+w\leq 0, and each v~\tilde{v}-trajectory is produced by freezing the coordinates ww and y1,,yn1y_{1},\ldots,y_{n-1}.

Since df~(v~)>0d\tilde{f}(\tilde{v})>0 in X~(F,ϵ)\tilde{X}(F,\epsilon), the field v~\tilde{v} is traversing in X~(F,ϵ)\tilde{X}(F,\epsilon). Hence each v~\tilde{v}-trajectory γX~(F,ϵ)\gamma\subset\tilde{X}(F,\epsilon) is either transversal to X^\partial\hat{X} at a pair of points, or it is simply tangent to X^\partial\hat{X} at a singleton, or does not intersect X^\hat{X}.

Consider the set ZX~(F,ϵ)Z\subset\tilde{X}(F,\epsilon) of v~\tilde{v}-trajectories through the points of X^\hat{X} (equivalently, the set of v~\tilde{v}-trajectories through 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v})). Every such trajectory γ\gamma is a closed oriented segment [a(γ),b(γ)][a(\gamma),b(\gamma)], where a(γ)b(γ)ZX~(F,ϵ)a(\gamma)\neq b(\gamma)\in\partial Z\subset\partial\tilde{X}(F,\epsilon). Thus, the variation varγ(f~)var_{\gamma}(\tilde{f}) of the function f~\tilde{f} along γ\gamma is strictly positive. Using compactness of X^\hat{X}, we get that infγZ{varγ(f~)}>0\inf_{\gamma\subset Z}\;\{var_{\gamma}(\tilde{f})\}>0.

Let Y=𝖽𝖾𝖿γZa(γ)ZY=_{\mathsf{def}}\bigcup_{\gamma\subset Z}a(\gamma)\subset\partial Z. Using the local model {u2+wϵ}\{u^{2}+w\leq\epsilon\}, we see that v~\tilde{v} is transversal to YY for all sufficiently small ϵ>0\epsilon>0.

Now, let us consider a new function on γ\gamma:

hγ(x)=𝖽𝖾𝖿(var[a(γ),x]f~)/(var[a(γ),b(γ)]f~).h_{\gamma}(x)=_{\mathsf{def}}(var_{[a(\gamma),\,x]}\tilde{f})\big{/}(var_{[a(\gamma),\,b(\gamma)]}\tilde{f}).

The function h:Zh:Z\to\mathbb{R}, defined as a collection of functions {hγ:γ}γZ\{h_{\gamma}:\gamma\to\mathbb{R}\}_{\gamma\subset Z}, is evidently a new Lyapunov function for v~\tilde{v} on ZZ. It is smooth thanks to the transversality of v~\tilde{v} to YY and the smooth dependence of solutions of ODEs on initial data. In fact, hh gives the product structure [0,1]×Y[0,1]\times Y to ZZ. Indeed, any point xZx\in Z is determined by the unique trajectory γx\gamma_{x} through xx and by the value at xx of the Lyapunov function hh of the v~\tilde{v}-flow.

Finally, the (v~)(-\tilde{v})-flow defines a smooth map p:1+X^(v^)Yp:\partial_{1}^{+}\hat{X}(\hat{v})\to Y which is a homeomorphism. In fact, 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}) is diffeomorphic to YY by a small perturbation of pp. ∎

Let α:XX^\alpha:X\to\hat{X} be a smooth submersion of a compact smooth manifold XX with boundary into the interior of a compact connected smooth manifold X^\hat{X} of the same dimension, where X^\partial\hat{X}\neq\emptyset. Assume that X^\hat{X} is equipped with a traversing vector field v^\hat{v} so that the pair (X^,v^)(\hat{X},\hat{v}) is convex in the sense of Definition 4.5. As before, we denote by v=α(v^)v=\alpha^{\dagger}(\hat{v}) the pull-back of v^\hat{v} under α\alpha. We denote by ff the pull-back α(f^)\alpha^{\ast}(\hat{f}) of the Lyapunov function f^\hat{f}. When α\alpha is a regular embedding, to simplify the notations, we identify XX and α(X)\alpha(X). Under this identification, v=v^|Xv=\hat{v}|_{X} and f=f^|Xf=\hat{f}|_{X}.

Definition 4.6.

Let (X^,v^)(\hat{X},\hat{v}) be a convex pair and let α:XX^\alpha:X\to\hat{X} be a submersion for which α\alpha^{\partial} is locally generic (in the sense of Definition 4.4) with respect to v^\hat{v}.

We say that α\alpha is kk-convex if k+X(α(v^))=\partial_{k}^{+}X(\alpha^{\dagger}(\hat{v}))=\emptyset, kk-concave if kX(α(v^))=\partial_{k}^{-}X(\alpha^{\dagger}(\hat{v}))=\emptyset, and kk-flat if kX(α(v^))=\partial_{k}X(\alpha^{\dagger}(\hat{v}))=\emptyset. \diamondsuit

Let a regular immersion α:XX^\alpha^{\partial}:\partial X\to\hat{X} be locally generic relative to v^\hat{v} in the sense of Definition 4.4. For each v^\hat{v}-trajectory γ^X^\hat{\gamma}\subset\hat{X}, we pick a point aγ^α(X)a\in\hat{\gamma}\cap\alpha(\partial X). Such aa belongs to the intersection of k=k(a)[1,n+1]k=k(a)\in[1,n+1] local branches of α(X)\alpha^{\partial}(\partial X), where k(a)=#((α)1(a))k(a)=\#((\alpha^{\partial})^{-1}(a)). In particular, as we remarked before, aa belongs to a unique collection of the α\alpha-images of the Morse strata {α(j(b)X(v))}bα1(a)\{\alpha(\partial_{j(b)}X(v))\}_{b\,\in\alpha^{-1}(a)} with the maximal possible indexes j(b)j(b) (recall that v:=α(v^)v:=\alpha^{\dagger}(\hat{v}) is the α\alpha-transfer of v^\hat{v}). The tangent spaces of these strata are in general position at aa, thanks to α\alpha^{\partial} being locally generic. Recall that this setting includes the cases where some or all the strata {α(j(b)X(v))}bα1(a)\{\alpha(\partial_{j(b)}X(v))\}_{b\,\in\alpha^{-1}(a)} are nn-dimensional, i.e., j(b)=1j(b)=1.

Let j(a)=𝖽𝖾𝖿bα1(a)j(b)k(a) and\displaystyle\text{ Let }\;j(a)=_{\mathsf{def}}\sum_{b\,\in\alpha^{-1}(a)}j(b)\;\geq\;k(a)\text{ \; and}
(4.3) j(a)(α(X))(v^)=𝖽𝖾𝖿bα1(a)α(j(b)X(v)),\displaystyle\partial_{j(a)}\big{(}\alpha(X)\big{)}(\hat{v})=_{\mathsf{def}}\bigcap_{\,b\,\in\alpha^{-1}(a)}\alpha\big{(}\partial_{j(b)}X(v)\big{)},

the latter equality being understood as an identity of the two germs at aa of the LHS and RHS loci. By Definition 4.4, the germ of j(a)(α(X))(v^)\partial_{j(a)}(\alpha(X))(\hat{v}) at aa is a smooth submanifold of X^\hat{X}, transversal to the trajectory γ^a\hat{\gamma}_{a}.

Let 𝖳aTaX^\mathsf{T}_{a}\subset T_{a}\hat{X} be the tangent space to the minimal stratum j(a)(α(X))\partial_{j(a)}(\alpha(X)) at aa (see (4.3)). With the help of the v^\hat{v}-flow, the subspace 𝖳aTaX^\mathsf{T}_{a}\subset T_{a}\hat{X} spreads to form a (dim(X^)j(a))(\dim(\hat{X})-j(a))-dimensional subbundle 𝖳a\mathsf{T}^{\bullet}_{a} of the tangent bundle TX^|γ^T\hat{X}\big{|}_{\hat{\gamma}} along the trajectory γ^\hat{\gamma}. We denote by 𝖳a\mathsf{T}^{\clubsuit}_{a} the image of 𝖳a\mathsf{T}^{\bullet}_{a} under the quotient map TX^|γ^TX^|γ^/Tγ^T\hat{X}|_{\hat{\gamma}}\to T\hat{X}|_{\hat{\gamma}}\big{/}T\hat{\gamma}, where Tγ^T\hat{\gamma} stands for the 11-bundle, tangent to γ^\hat{\gamma}.

We introduce a sightly modified version of Definition 3.2 from [K2], a modification that applies to submersions α:XX^\alpha:X\to\hat{X}.

Definition 4.7.

Let v^\hat{v} be a traversing vector field on a compact connected smooth manifold on X^\hat{X} with boundary. We say that a submersion α:Xint(X^)\alpha:X\to\textup{int}(\hat{X}) is traversally generic relative to v^\hat{v}, if:

  • α\alpha^{\partial} is locally generic in the sense of Definition 4.4 with respect to v^\hat{v},

  • for each v^\hat{v}-trajectory γ^X^\hat{\gamma}\subset\hat{X}, the subbundles {𝖳a}aγ^α(X)\{\mathsf{T}^{\clubsuit}_{a}\}_{a\,\in\,\hat{\gamma}\,\cap\,\alpha(\partial X)} are in general position in the normal to γ^\hat{\gamma} (trivial) nn-bundle (TX^|γ^)/Tγ^(T\hat{X}|_{\hat{\gamma}})\big{/}T\hat{\gamma}. \diamondsuit

Example 4.4.

The patterns in Fig.2 may be stacked vertically along a trajectory γ^\hat{\gamma}. To get a traversally generic piles in the vicinity of γ\gamma, we obey the following rules: (1) to any stack, we may add any number of configuration of type aa from Fig.1, as long as the prescribed parity of m(γ^)0mod2m(\hat{\gamma})\equiv 0\mod 2 is not violated; (2) no two configurations of multiplicity 33 (of the types c,e,fc,e,f) reside on γ^\hat{\gamma}; (3) at most two configurations of multiplicity 22 (of the types b,db,d) reside on γ^\hat{\gamma}, moreover, the (v^)(-\hat{v})-directed projections on the transversal section SS to γ^\hat{\gamma} of the fold loci (as in dd) or/and of the simple self-intersections (as in bb) must be transversal in SS. \diamondsuit

Let α\alpha be traversally generic relative to v^\hat{v}. Then, for any v^\hat{v}-trajectory γ^\hat{\gamma}, by counting the dimensions of the bundles {𝖳a}aγ^α(X)\{\mathsf{T}^{\clubsuit}_{a}\}_{a\,\in\,\hat{\gamma}\,\cap\,\alpha(\partial X)} and using that they are in general position in the nn-dimensional bundle, normal to γ^\hat{\gamma}, we get that the reduced multiplicity

(4.4) m(γ^)=𝖽𝖾𝖿aγ^α(X)(j(a)1)n.\displaystyle m^{\prime}(\hat{\gamma})=_{\mathsf{def}}\sum_{a\,\in\,\hat{\gamma}\,\cap\,\alpha(\partial X)}\big{(}j(a)-1\big{)}\,\leq\,n.

For a traversally generic α\alpha and a trajectory γ^\hat{\gamma}, let γX\gamma\subset X be any segment in α1(γ^)X\alpha^{-1}(\hat{\gamma})\subset X which is bounded by a pair of points in X\partial X. Then, by Theorem 3.5 from [K2], the total multiplicity m(γ)2n+2m(\gamma)\leq 2n+2. Although, for a traversally generic α\alpha, there is no α\alpha-independent constraint on the total multiplicity m(γ^)=𝖽𝖾𝖿xγ^α(X)j(x)m(\hat{\gamma})=_{\mathsf{def}}\sum_{x\,\in\,\hat{\gamma}\,\cap\,\alpha(\partial X)}j(x), there is an universal constraint on the cardinality of the subset γ^{2}γ^α(X)\hat{\gamma}_{\{\geq 2\}}\subset\hat{\gamma}\cap\alpha(\partial X), consisting of points xx whose multiplicity m(x)2m(x)\geq 2. Namely, #(γ^{2})n\#(\hat{\gamma}_{\{\geq 2\}})\leq n, since no more than nn proper vector subspaces may be in general position in an ambient nn-dimensional vector space.

Now we are ready to introduce the central notion of a convex pseudo-envelop.

Definition 4.8.

Let v^\hat{v} be a traversing vector field on a compact connected smooth manifold X^\hat{X} with boundary. We assume that (X^,v^)(\hat{X},\hat{v}) is convex in the sense of Definition 4.5.

We call such a pair (X^,v^)(\hat{X},\hat{v}) a convex pseudo-envelop of a submersion α:Xint(X^)\alpha:X\to\textup{int}(\hat{X}), if α\alpha^{\partial} is locally generic relative to v^\hat{v}. We think of XX as being equipped with the pull-back vector field v=α(v^)v=\alpha^{\dagger}(\hat{v}), so that (X,v)(X,v) is “enveloped” by (X^,v^)(\hat{X},\hat{v}).

If α\alpha is a locally generic regular embedding, then we call (X^,v^)(\hat{X},\hat{v}) a convex envelop of α\alpha.

\diamondsuit

Remark 4.2.

Not all traversally generic pairs (X,v)(X,v) admit convex envelops (X^,v^)(\hat{X},\hat{v}). For example, if XX has two closed submanifolds (or singular cycles), MM and NN, of complementary dimensions with a nonzero algebraic intersection number MNM\circ N, then no convex envelop (X^,v^)(\hat{X},\hat{v}) of (X,v)(X,v) exists. Indeed, with the help of the (v^)(-\hat{v})-flow, MM is cobordant in X^\hat{X} to a cycle MM^{\prime} which resides in 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}). Similarly, with the help of the v^\hat{v}-flow, NN is cobordant in X^\hat{X} to a cycle NN^{\prime} which resides in 1X^(v^)\partial_{1}^{-}\hat{X}(\hat{v}). Note that if MN0M\circ N\neq 0 in XX, then evidently the same property holds in any X^X\hat{X}\supset X. Since MM^{\prime} and NN^{\prime} are disjoint cycles, their intersection number MN=0M^{\prime}\circ N^{\prime}=0, which contradicts to the assumption MN0M\circ N\neq 0. In particular, the non-triviality of 𝗌𝗂𝗀𝗇(X)\mathsf{sign}(X), the Wall relative signature of XX [W], obstructs the existence of a convex envelop for any traversing vv on XX. As Fig.1 testifies, this argument does not rule out the existence of a convex pseudo-envelop for (X,v)(X,v). In fact, Fig.1 shows that any compact oriented surface XX can be enveloped. \diamondsuit

We denote by 𝖲𝗎𝖻(X,X^)\mathsf{Sub}(X,\hat{X}) the space of smooth submersions α:X𝗂𝗇𝗍(X^)\alpha:X\to\mathsf{int}(\hat{X}).

When α\alpha is an embedding, by Theorem 3.5 from [K2], in the space 𝒱𝗍𝗋𝖺𝗏(X^,α)\mathcal{V}_{\mathsf{trav}}(\hat{X},\alpha) of traversing vector fields v^\hat{v} on X^\hat{X}, there is an open and dense subset 𝒱(X^,α)\mathcal{V}^{\ddagger}(\hat{X},\alpha) such that α\alpha is traversally generic (see Definition 4.7) with respect to v^𝒱(X^,α)\hat{v}\in\mathcal{V}^{\ddagger}(\hat{X},\alpha).

On the other hand, the property of a submersion α\alpha to be traversally generic with respect to a given traversing vector field v^\hat{v} (see Definition 4.7) is an “open” property in the CC^{\infty}-topology on 𝖲𝗎𝖻(X,X^)\mathsf{Sub}(X,\hat{X}), since it may be expressed in terms of mutual transversality of the relevant strata in the appropriate jet spaces.

We conjecture that the transversal generality of submersions α\alpha with respect to a fixed convex pair (X^,v^)(\hat{X},\hat{v}) is also a “dense” property. Among other things, the next Theorem 4.1 shows that this conjecture is valid for the regular embeddings α\alpha which admit convex envelops. However, for general submersions α\alpha, we are able only to prove that a somewhat weaker property “is dense”. That property is described in the second claim of Theorem 4.1. Speaking informally, we can insure by α\alpha-perturbations the general positions of the singularities of the maps α:XX^\alpha^{\partial}:\partial X\to\hat{X} and πα:X1+X^(v^)\pi\circ\alpha^{\partial}:\partial X\to\partial_{1}^{+}\hat{X}(\hat{v}) separately, but not mutually. Here the map π:X^1+X^(v^)\pi:\hat{X}\to\partial_{1}^{+}\hat{X}(\hat{v}) is defined by the (v^)(-\hat{v})-directed convex flow.

Theorem 4.1.

Let (X^,v^)(\hat{X},\hat{v}) be a convex pair, and XX a compact smooth manifold with boundary, dim(X)=dim(X^)=n+1\dim(X)=\dim(\hat{X})=n+1. Assume that 𝖲𝗎𝖻(X,X^)\mathsf{Sub}(X,\hat{X})\neq\emptyset.

  • There is an open and dense subset 𝒪𝖲𝗎𝖻(X,X^)\mathcal{O}\subset\mathsf{Sub}(X,\hat{X}) such that, for any α𝒪\alpha\in\mathcal{O}, the local branches of α(X)\alpha(\partial X) are in general position in TaX^T_{a}\hat{X} at every point aα(X)a\in\alpha(\partial X).

  • For any α𝒪\alpha\in\mathcal{O} and for each v^\hat{v}-trajectory γ^X^\hat{\gamma}\subset\hat{X}, the v^\hat{v}-invariant subbundles {𝖳a}aγ^α(X)\{\mathsf{T}^{\clubsuit}_{a}\}_{a\,\in\;\hat{\gamma}\,\cap\,\alpha(\partial X)}, generated by the intersections

    b(α)1(a),j(b)2α[Tb(j(b)X(v))]\bigcap_{b\in(\alpha^{\partial})^{-1}(a),\;j(b)\geq 2}\alpha_{\ast}\big{[}T_{b}\big{(}\partial_{j(b)}X(v)\big{)}\big{]}

    of the tangent spaces to the Morse strata {j(b)X(v)}b(α)1(a),j(b)2\big{\{}\partial_{j(b)}X(v)\big{\}}_{b\in(\alpha^{\partial})^{-1}(a),\;j(b)\geq 2}, are in general position in the normal to γ^\hat{\gamma} (trivial) nn-bundle (TX^|γ^)/Tγ^(T\hat{X}|_{\hat{\gamma}})\big{/}T\hat{\gamma}. Here vv is the pull-back of v^\hat{v} under α\alpha.

Proof.

We assume that the openness of 𝒪\mathcal{O} in 𝖲𝗎𝖻(X,X^)\mathsf{Sub}(X,\hat{X}) is clear, due to the compactness of XX, and will present the arguments that validate the density of 𝒪\mathcal{O} in 𝖲𝗎𝖻(X,X^)\mathsf{Sub}(X,\hat{X}). We divide the proof into three steps, marked as (i), (ii), and (iii).

(i) By [LS], the set 𝒩C(X,X^)\mathcal{N}\subset C^{\infty}(\partial X,\hat{X}) of smooth maps β:XX^\beta:\partial X\to\hat{X}, such that β\beta is a kk-normal immersion in the sense of [LS] (see also Definition 3.3 from [K9]) for all kn+1k\leq n+1, is open and dense.

First we aim to show that, for a given submersion α𝖲𝗎𝖻(X,X^)\alpha\in\mathsf{Sub}(X,\hat{X}), there is an open set 𝒪α𝖲𝗎𝖻(X,X^)\mathcal{O}_{\alpha}\subset\mathsf{Sub}(X,\hat{X}) such that α𝖼𝗅𝗈𝗌𝗎𝗋𝖾(𝒪α)\alpha\in\mathsf{closure}(\mathcal{O}_{\alpha}) and, for any α~𝒪α\tilde{\alpha}\in\mathcal{O}_{\alpha}, the submersion α~𝒩\tilde{\alpha}^{\partial}\in\mathcal{N}.

With this goal in mind, we choose an auxiliary metric gg on X^\hat{X} such that: (1)(1) the boundary X^\partial\hat{X} is convex in gg, and (2)(2) there is ϵ0>0\epsilon_{0}>0 such that any two points x,yX^x,y\in\hat{X} that are less than ϵ0\epsilon_{0}-apart are connected by a single geodesic arc. Using the submersion α\alpha, we pull-back gg to a Riemannian metric gαg^{\dagger}_{\alpha} on XX.

For some ϵ<ϵ0\epsilon<\epsilon_{0}, the ϵ\epsilon-neighborhood CϵXC_{\epsilon}\subset X of X\partial X in the metric gαg^{\dagger}_{\alpha} has a product structure ψ:X×[0,ϵ]Cϵ\psi:\partial X\times[0,\epsilon]\stackrel{{\scriptstyle\approx}}{{\to}}C_{\epsilon}, so that the curve δx=𝖽𝖾𝖿ψ(x×[0,ϵ])\delta_{x}=_{\mathsf{def}}\psi(x\times[0,\epsilon]) is the unique geodesic in gαg^{\dagger}_{\alpha}, normal to X\partial X at xXx\in\partial X. Using the diffeomorphism ψ\psi, we construct a smooth diffeomotopy {ψt:XX}t[0,1]\{\psi_{t}:X\to X\}_{t\in[0,1]} (via the flow inward normal to X\partial X) so that ψ0=𝗂𝖽X\psi_{0}=\mathsf{id}_{X} and ψ1(X)=X𝗂𝗇𝗍(Cϵ)\psi_{1}(X)=X\setminus\mathsf{int}(C_{\epsilon}).

Using the convexity of X^\partial\hat{X} in gg and the choice of ϵ<ϵ0\epsilon<\epsilon_{0} we get the following claim: for any β\beta that approximates α\alpha^{\partial} and each xXx\in\partial X, there exists the unique geodesic δ^xX^\hat{\delta}_{x}\subset\hat{X} in the metric gg that connects the points β(x)\beta(x) and αψ1(x)\alpha\circ\psi_{1}(x).

We pick such an approximation β𝒩\beta\in\mathcal{N} of α\alpha^{\partial}. Let a smooth map α(β,ϵ):XX^\alpha^{\sharp}(\beta,\epsilon):X\to\hat{X} be defined by the two properties: (1) α(β,ϵ)|XCϵ:=αψ1\alpha^{\sharp}(\beta,\epsilon)|_{X\setminus C_{\epsilon}}:=\alpha\circ\psi_{1}, (2) for each xXx\in\partial X, the diffeomorphism α(β,ϵ)|:δxδ^x\alpha^{\sharp}(\beta,\epsilon)|:\delta_{x}\to\hat{\delta}_{x} is an isometry with respect to gαg^{\dagger}_{\alpha} and gg along the two geodesic arcs. By picking ϵ\epsilon small enough and β𝒩\beta\in\mathcal{N} sufficiently CC^{\infty}-close to α\alpha^{\partial}, we get that α(β,ϵ)\alpha^{\sharp}(\beta,\epsilon) is CC^{\infty}-close to α\alpha. Therefore, we may assume that α(β,ϵ)𝖲𝗎𝖻(X,X^)\alpha^{\sharp}(\beta,\epsilon)\in\mathsf{Sub}(X,\hat{X}) for an appropriate choice of β\beta that approximates α\alpha^{\partial}. By its construction, (α(β,ϵ))𝒩(\alpha^{\sharp}(\beta,\epsilon))^{\partial}\in\mathcal{N}.

(ii) For a given smooth map β:X𝗂𝗇𝗍(1+X^(v^))\beta:\partial X\to\mathsf{int}(\partial_{1}^{+}\hat{X}(\hat{v})), let β\beta_{\ast} denote the differential of β\beta.

 Let 𝒮(1)(β)=𝖽𝖾𝖿{xX|𝗋𝗄(β)n1}.\text{ Let \;}\mathcal{S}^{(1)}(\beta)=_{\mathsf{def}}\big{\{}x\in\partial X\big{|}\;\mathsf{rk}(\beta_{\ast})\leq n-1\big{\}}.

If S(1)(β)S^{(1)}(\beta) is a smooth manifold, then the locus

𝒮(1,1)(β)=𝖽𝖾𝖿{x𝒮(1)(β)|𝗋𝗄(β|𝒮(1)(β))n2}\mathcal{S}^{(1,1)}(\beta)=_{\mathsf{def}}\big{\{}x\in\mathcal{S}^{(1)}(\beta)\big{|}\;\mathsf{rk}(\beta_{\ast}|_{\mathcal{S}^{(1)}(\beta)})\leq n-2\big{\}}

is well-defined. Continuing this way, the filtration of X\partial X by the loci {𝒮[k](β)=𝖽𝖾𝖿𝒮ω(β)}k\big{\{}\mathcal{S}_{[k]}(\beta)=_{\mathsf{def}}\mathcal{S}^{\omega}(\beta)\big{\}}_{k}, where kn+1k\leq n+1 and ω=(1,1,,1k)\omega=(\underbrace{1,1,\ldots,1}_{k}) is introduced. Applying Boardman’s Theorems 15.1-15.3, [Bo], the subset \mathcal{B} of maps βC(X,1+X^(v^))\beta\in C^{\infty}(\partial X,\partial_{1}^{+}\hat{X}(\hat{v})), for which all the strata {𝒮[k]=𝖽𝖾𝖿𝒮[k](β)}\{\mathcal{S}_{[k]}=_{\mathsf{def}}\mathcal{S}_{[k]}(\beta)\} are smooth manifolds and all the maps

{β|:𝒮[k]=𝖽𝖾𝖿𝒮[k]𝒮[k+1]𝗂𝗇𝗍(1+X^(v^))}k\big{\{}\beta|:\mathcal{S}_{[k]}^{\circ}=_{\mathsf{def}}\mathcal{S}_{[k]}\setminus\mathcal{S}_{[k+1]}\to\mathsf{int}(\partial_{1}^{+}\hat{X}(\hat{v}))\big{\}}_{k}

are immersions, is open (X\partial X is compact) and dense. Note that each xXx\in\partial X belongs to a unique pure stratum 𝒮[k(x)]\mathcal{S}_{[k(x)]}^{\circ}. Moreover, using the Thom Multijet Transversality Theorem (see [GG], Theorem 4.13), the subspace of \mathcal{B}, formed maps β\beta for which all the spaces {β(Tx𝒮[k(x)])}x(πβ)1(y)\{\beta_{\ast}(T_{x}\mathcal{S}_{[k(x)]})\}_{x\in(\pi\circ\beta)^{-1}(y)} are in general position in Ty1+X^(v^)T_{y}\partial_{1}^{+}\hat{X}(\hat{v}) for all k(x)2k(x)\geq 2 and all y𝗂𝗇𝗍(1+X^(v^))y\in\mathsf{int}(\partial_{1}^{+}\hat{X}(\hat{v})), form an open and dense subset 𝖭𝖢\mathcal{B}_{\mathsf{NC}}\subset\mathcal{B} (see [GG], Theorem 5.2). Here “𝖭𝖢\mathsf{NC}” abbreviates the condition known as “normal crossings”. Note that if k(x)=1k(x)=1, then β(TxS[k(x)])\beta_{\ast}(T_{x}S_{[k(x)]}) coincides with Tβ(x)1+X^(v^)T_{\beta(x)}\partial_{1}^{+}\hat{X}(\hat{v}).

We stress that, applying the previous arguments to the map β=πα\beta=\pi\circ\alpha^{\partial}, the strata bα1(a)α(j(b)X(v))\bigcap_{\,b\,\in\alpha^{-1}(a)}\alpha\big{(}\partial_{j(b)}X(v)\big{)} that involve bb’s with j(b)=1j(b)=1 become “β\beta-invisible”, thanks to the “erasing” action of π\pi_{\ast} on all nn-dimensional spaces α(Ta(1X(v)))\alpha\big{(}T_{a}(\partial_{1}X(v))\big{)}. Because of this short comming, we cannot claim that the traversally generic α\alpha to v^\hat{v} (see Definition 4.7) are dense in 𝖲𝗎𝖻(X,X^)\mathsf{Sub}(X,\hat{X}).

(iii) Note that the map π:X^𝗂𝗇𝗍(1+X^(v^))\pi:\hat{X}\to\mathsf{int}(\partial_{1}^{+}\hat{X}(\hat{v})), defined by the (v^)(-\hat{v})-flow, is smooth due to v^\hat{v} being convex; moreover, its restriction to 𝗂𝗇𝗍(X^)\mathsf{int}(\hat{X}) is a submersion.

For a given α𝖲𝗎𝖻(X,X^)\alpha\in\mathsf{Sub}(X,\hat{X}), we form the composition α=𝖽𝖾𝖿πα\alpha^{\flat}=_{\mathsf{def}}\pi\circ\alpha^{\partial}.

By (i), we can approximate α\alpha by a new submersion α1=α(β,ϵ)\alpha_{1}=\alpha^{\sharp}(\beta,\epsilon) such that α1𝒩\alpha_{1}^{\partial}\in\mathcal{N}.

By (ii), we can approximate α1=𝖽𝖾𝖿πα1\alpha_{1}^{\flat}=_{\mathsf{def}}\pi\circ\alpha_{1}^{\partial} by a smooth map β1:X𝗂𝗇𝗍(1+X^(v^))\beta_{1}^{\flat}:\partial X\to\mathsf{int}(\partial_{1}^{+}\hat{X}(\hat{v})) such that β1𝖭𝖢\beta_{1}^{\flat}\in\mathcal{B}_{\mathsf{NC}}.

Let us fix a Lyapunov function f^:X^\hat{f}:\hat{X}\to\mathbb{R} for v^\hat{v}. In the spirit of Lemma 4.3, the Lyapunov function f^\hat{f} for v^\hat{v} and the projection π:X^1+X^(v^)\pi:\hat{X}\to\partial_{1}^{+}\hat{X}(\hat{v}) define global smooth “coordinates” in the interior of X^\hat{X}. That is, each pair (t,y)(t,y), where y1+X^(v^)y\in\partial_{1}^{+}\hat{X}(\hat{v}) and tf^(π1(y))t\in\hat{f}(\pi^{-1}(y)), determines a unique point xX^x\in\hat{X} such that f^(x)=t\hat{f}(x)=t and π(x)=y\pi(x)=y. Let f1=α1(f^)f_{1}=\alpha^{\ast}_{1}(\hat{f}).

For the map β1𝖭𝖢\beta_{1}^{\flat}\in\mathcal{B}_{\mathsf{NC}}, we define a smooth map β1=𝖽𝖾𝖿β1(β1,f^):XX^\beta_{1}=_{\mathsf{def}}\beta_{1}(\beta_{1}^{\flat},\hat{f}):\partial X\to\hat{X} by the formula β1(x)=y\beta_{1}(x)=y, where xXx\in\partial X and yy is the unique point on the v^\hat{v}-trajectory γ^\hat{\gamma} through β1(x)1+X^(v^)\beta_{1}^{\flat}(x)\in\partial_{1}^{+}\hat{X}(\hat{v}) such that f^(y)=f1(x)\hat{f}(y)=f_{1}(x). Note that by choosing β1\beta_{1}^{\flat} sufficiently close to α1\alpha_{1}^{\flat}, we insure that β1\beta_{1} is sufficiently close to α1\alpha_{1}^{\partial}. Thus we may assume that β1𝒩\beta_{1}\in\mathcal{N}.

Recycling the argument that revolves around the construction of α(β,ϵ):XX^\alpha^{\sharp}(\beta,\epsilon):X\to\hat{X} in part (i) of the proof, we form the submersion α2=𝖽𝖾𝖿α1(β1,ϵ):XX^\alpha_{2}=_{\mathsf{def}}\alpha_{1}^{\sharp}(\beta_{1},\epsilon):X\to\hat{X}. By the construction of α2\alpha_{2}, we have α2=β1𝖭𝖢\alpha_{2}^{\flat}=\beta_{1}^{\flat}\in\mathcal{B}_{\mathsf{NC}} and β1𝒩\beta_{1}\in\mathcal{N}.

Therefore, we have shown that any given submersion α\alpha admits an approximation by some α2𝖲𝗎𝖻(X,X^)\alpha_{2}\in\mathsf{Sub}(X,\hat{X}) such that α2𝒩\alpha_{2}^{\partial}\in\mathcal{N} and α2𝖭𝖢\alpha_{2}^{\flat}\in\mathcal{B}_{\mathsf{NC}}. These are exactly the two properties that describe the space 𝒪\mathcal{O} in the theorem. ∎

Corollary 4.2.

Let (X^,v^)(\hat{X},\hat{v}) be a convex pair, and XX a compact smooth manifold with boundary, dim(X)=dim(X^)=n+1\dim(X)=\dim(\hat{X})=n+1.

The regular embeddings α:XX^\alpha:X\subset\hat{X} that are traversally generic with respect to v^\hat{v} (see Definition 4.7) form an open and dense set in the space of all regular smooth embeddings.

Proof.

Since, for a regular embedding α:XX^\alpha:X\hookrightarrow\hat{X}, α\alpha^{\partial} is an embedding, the first claim of Theorem 4.1 is vacuous, and the second claim insures that α\alpha is traversally generic. ∎

Remark 4.3.

Consider the kk-multiple self-intersection manifolds {Σkα}k\{\Sigma_{k}^{\alpha^{\partial}}\}_{k} of kk-normal (see [LS]) immersions α:XX^\alpha^{\partial}:\partial X\to\hat{X}. By definition, Σkα\Sigma_{k}^{\alpha^{\partial}} is a submanifold of the kk-fold product (X)k(\partial X)^{k}, the preimage of the diagonal Δ(X^)k\Delta\subset(\hat{X})^{k} under the transversal to it map (α)k(\alpha^{\partial})^{k}. The projection p1p_{1} of Σkα\Sigma_{k}^{\alpha^{\partial}} on the first factor X\partial X of the product (X)k(\partial X)^{k} is an immersion [LS]. By composing p1p_{1} with α\alpha^{\partial}, we get an immersion of αp1:ΣkαX^\alpha^{\partial}\circ p_{1}:\Sigma_{k}^{\alpha^{\partial}}\to\hat{X}. By using the convex (v^)(-\hat{v})-flow, we get a map π:X^1+X^(v^)\pi:\hat{X}\to\partial_{1}^{+}\hat{X}(\hat{v}). Finally, we obtain a smooth composite map παp1:Σkα1+X^(v^)\pi\circ\alpha^{\partial}\circ p_{1}:\Sigma_{k}^{\alpha^{\partial}}\to\partial_{1}^{+}\hat{X}(\hat{v}).

We notice that, under the hypotheses and notations of Theorem 4.1, if v^\hat{v} is tangent at aa to the intersection b(α)1(a)α(j(b)X(v))\bigcap_{b\in(\alpha^{\partial})^{-1}(a)}\;\alpha_{\ast}\big{(}\partial_{j(b)}X(v)\big{)}, then, evidently, each local branch α(j(b)X(v))\alpha_{\ast}(\partial_{j(b_{\star})}X(v)), b(α)1(a),b_{\star}\in(\alpha^{\partial})^{-1}(a), is tangent to v^\hat{v} at aa. Therefore, j(b)2j(b_{\star})\geq 2. In other words, for k2k\geq 2, the singular locus of the map παp1:Σkα1+X^(v^)\pi\circ\alpha^{\partial}\circ p_{1}:\Sigma_{k}^{\alpha^{\partial}}\to\partial_{1}^{+}\hat{X}(\hat{v}) is always contained in the singular locus of the map παp1:(X)k1+X^(v^)\pi\circ\alpha^{\partial}\circ p_{1}:(\partial X)^{k}\to\partial_{1}^{+}\hat{X}(\hat{v}). \diamondsuit

4.4. Quasitopies of convex envelops and pseudo-envelops

Now, let us modify Definition 3.7 from [K9] and Definition 3.1 from this paper, so that they apply to convex pseudo-envelops of traversing flows (see Fig. 3). This modification is central to our efforts.

Definition 4.9.

Fix natural even numbers ddd\leq d^{\prime} and consider a closed subposets ΘΘ\Theta^{\prime}\subset\Theta of the universal poset 𝛀\mathbf{\Omega} from Section 2, such that ()Θ(\emptyset)\notin\Theta .

Let X^\hat{X} be a (n+1)(n+1)-dimensional compact manifold and v^\hat{v} a convex traversing vector field on it. Let Z^=𝖽𝖾𝖿X^×[0,1]\hat{Z}=_{\mathsf{def}}\hat{X}\times[0,1]. We denote by v^\hat{v}^{\bullet} the vector field on Z^\hat{Z} that is tangent to each slice X^×{t}\hat{X}\times\{t\}, t[0,1]t\in[0,1], and is equal to v^\hat{v} there.

We say that a two convex pseudo-envelops, α0:X0X^\alpha_{0}:X_{0}\to\hat{X} and α1:X1X^\alpha_{1}:X_{1}\to\hat{X}, are (d,d;𝐜Θ,𝐜Θ)(d,d^{\prime};\mathbf{c}\Theta,\mathbf{c}\Theta^{\prime})-quasitopic in X^\hat{X}, if there exists a compact smooth orientable (n+2)(n+2)-manifold WW,444with corners X0X1\partial X_{0}\coprod\partial X_{1} whose boundary W=(X0X1){X0X1}δW\partial W=(X_{0}\coprod X_{1})\bigcup_{\{\partial X_{0}\coprod\partial X_{1}\}}\delta W, and a smooth submersion A:WZ^A:W\to\hat{Z} so that:

  • A|X0=α0A|_{X_{0}}=\alpha_{0} and A|X1=α1A|_{X_{1}}=\alpha_{1};

  • for each zX^×[0,1]z\in\hat{X}\times\,\partial[0,1], the total multiplicity mA(γ^z)m_{A}(\hat{\gamma}_{z}) of the v^\hat{v}^{\bullet}-trajectory γ^z\hat{\gamma}_{z} through zz, relatively to A(X0X1)A(\partial X_{0}\coprod\partial X_{1}), satisfies the constraints mA(γ^z)dm_{A}(\hat{\gamma}_{z})\leq d, mA(γ^z)0mod2m_{A}(\hat{\gamma}_{z})\equiv 0\mod 2, and the combinatorial tangency pattern ωA(γ^z)\omega^{A}(\hat{\gamma}_{z}) of γ^z\hat{\gamma}_{z} with respect to A(X0X1)A(\partial X_{0}\coprod\partial X_{1}) does not belong to Θ\Theta;

  • for each zZ^z\in\hat{Z}, the total multiplicity mA(γ^z)m_{A}(\hat{\gamma}_{z}) of γ^z\hat{\gamma}_{z} with respect to A(δW)A(\delta W) satisfies the constraints mA(γ^z)dm_{A}(\hat{\gamma}_{z})\leq d^{\prime}, mA(γ^z)0mod2m_{A}(\hat{\gamma}_{z})\equiv 0\mod 2, and the combinatorial tangency pattern ωA(γ^z)\omega^{A}(\hat{\gamma}_{z}) of γ^z\hat{\gamma}_{z} with respect to A(δW)A(\delta W) does not belong to Θ\Theta^{\prime};

We denote by 𝒬𝒯d,d𝗌𝗎𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}_{d,d^{\prime}}^{\mathsf{sub}}(\hat{X},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) the set of quasitopy classes of such convex pseudo-envelops α:(X,α(v^))(X^,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(\hat{X},\hat{v}).

If we insist that α0,α1\alpha_{0},\alpha_{1}, and AA are embeddings, then we get 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}_{d,d^{\prime}}^{\mathsf{emb}}(\hat{X},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}), the set of quasitopy classes of convex envelops. \diamondsuit

It is easy to check that the quasitopy of convex pseudo-envelops (convex envelops) α:(X,α(v^))(X^,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(\hat{X},\hat{v}) is an equivalence relation.

Refer to caption


Figure 3. The ingredients of Definition 4.9. For simplicity of the depiction, the maps α0,α1,A\alpha_{0},\alpha_{1},A are shown as embeddings.

We are in position to state one of the main results of this paper.

Theorem 4.2.

Let ΘΘ𝛀\Theta^{\prime}\subset\Theta\subset\mathbf{\Omega} be closed subposets that do not contain the element ()(\emptyset). Assume that ddd^{\prime}\geq d and dd0mod2d^{\prime}\equiv d\equiv 0\mod 2. Let X^\hat{X} be a smooth compact connected (n+1)(n+1)-dimensional manifold, equipped with a convex traversing vector field v^\hat{v}.

Then, under the notations of Definition 3.2, there is a canonical bijection

Φ𝖾𝗆𝖻:𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\displaystyle\Phi^{\mathsf{emb}}:\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}
(4.5) [[(1+X^(v^),2X^(v^)),ϵd,d:(𝒫d𝐜Θ,pt)(𝒫d𝐜Θ,pt)]]\displaystyle\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\big{[}\big{[}(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v})),\;\epsilon_{d,d^{\prime}}:(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt)\to(\mathcal{P}_{d^{\prime}}^{\mathbf{c}\Theta^{\prime}},pt^{\prime})\big{]}\big{]}

and a canonical surjection

Φ𝗌𝗎𝖻:𝒬𝒯d,d𝗌𝗎𝖻(X^,v^;𝐜Θ;𝐜Θ)epi\displaystyle\Phi^{\mathsf{sub}}:\mathcal{QT}^{\mathsf{sub}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\stackrel{{\scriptstyle epi}}{{\longrightarrow}}
(4.6) epi[[(1+X^(v^),2X^(v^)),ϵd,d:(𝒫d𝐜Θ,pt)(𝒫d𝐜Θ,pt)]].\displaystyle\stackrel{{\scriptstyle epi}}{{\longrightarrow}}\big{[}\big{[}(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v})),\;\epsilon_{d,d^{\prime}}:(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt)\to(\mathcal{P}_{d^{\prime}}^{\mathbf{c}\Theta^{\prime}},pt^{\prime})\big{]}\big{]}.

The map Φ𝗌𝗎𝖻\Phi^{\mathsf{sub}} admits a right inverse.

Proof.

By Lemma 4.3, any convex pseudo-envelop α:XX^\alpha:X\to\hat{X} may be incapsulated into a convex pseudo-envelop α~:XX^[0,1]×Y\tilde{\alpha}:X\to\hat{X}\subset[0,1]\times Y of the product type (we called a capsule [0,1]×Y[0,1]\times Ya box”), where YY is homeomorphic to 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}), and its boundary Y\partial Y to 2X^(v^)\partial_{2}^{-}\hat{X}(\hat{v}).

Conversely, any submersion α~:X[0,1]×Y\tilde{\alpha}:X\to[0,1]\times Y, where YY is a smooth compact nn-manifold, by rounding the corners [0,1]×Y\partial[0,1]\times\partial Y of [0,1]×Y[0,1]\times Y produces a convex pseudo-envelop (X^,v^)([0,1]×Y,v~)(\hat{X},\hat{v})\subset([0,1]\times Y,\tilde{v}), where v~0\tilde{v}\neq 0 is tangent to the fibers of [0,1]×YY[0,1]\times Y\to Y. Similarly, the cobordisms AA between pairs of quasitopies α0,α1\alpha_{0},\alpha_{1} can be incapsulated in boxes of the form ([0,1]×Y)×[0,1]([0,1]\times Y)\times[0,1]. Therefore, all the results from Section 3 in [K9] apply to α~\tilde{\alpha}^{\partial}; in particular, the pivotal Theorem 3.2 from [K9] applies. With its help, the maps Φ𝗌𝗎𝖻\Phi^{\mathsf{sub}} from (4.2) and Φ𝖾𝗆𝖻\Phi^{\mathsf{emb}} from (4.2) are generated as compositions of maps p(v~):(1+X^(v^),2X^(v^))(Y,Y)p(\tilde{v}):(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v}))\to(Y,\partial Y) with the maps {Φα~:(Y,Y)(𝒫d𝐜Θ,𝒫d())}\{\Phi^{\tilde{\alpha}^{\partial}}:(Y,\partial Y)\to\mathcal{(}\mathcal{P}_{d}^{\mathbf{c}\Theta},\mathcal{P}_{d}^{(\emptyset)})\} from Theorem 3.2 in [K9]; the latter map being generated by the locus α~(X)×Y\tilde{\alpha}(\partial X)\subset\mathbb{R}\times Y. Note that the cell 𝒫d()𝒫d\mathcal{P}_{d}^{(\emptyset)}\subset\mathcal{P}_{d} contracts to a singleton pt𝒫d()pt\in\mathcal{P}_{d}^{(\emptyset)}.

Similarly, for any quasitopy A:WX^×[0,1]A:W\to\hat{X}\times[0,1] as in Definition 4.9, the product Y×[0,1]Y\times[0,1] that encapsulates X^×[0,1]\hat{X}\times[0,1] is mapped to 𝒫d𝐜Θ\mathcal{P}_{d^{\prime}}^{\mathbf{c}\Theta^{\prime}} with the help of A(δW)A(\delta W), while Y×[0,1]\partial Y\times[0,1] are mapped to the cell 𝒫d()\mathcal{P}_{d^{\prime}}^{(\emptyset)}, which contracts to the singleton ϵd,d(pt)𝒫d()\epsilon_{d,d^{\prime}}(pt)\in\mathcal{P}_{d^{\prime}}^{(\emptyset)}.

As in Theorem 3.2 from [K9], and by similar transversality arguments, Φ𝖾𝗆𝖻\Phi^{\mathsf{emb}} is a 1-to-1 map. Here we need to use Lemma 3.4 from [K9] to conclude that, under a d\partial\mathcal{E}_{d}-regular map (see Definition 3.4 from [K9]), the preimage of the hypersurface d×𝒫d\partial\mathcal{E}_{d}\subset\mathbb{R}\times\mathcal{P}_{d} (see (2)) bounds a compact manifold XX in X^\hat{X}, provided d0mod2d\equiv 0\mod 2. Again, as in Proposition 3.5 from [K9], the bijective map in (4.2) helps to prove that the map in (4.2) is a split surjective one. ∎

Corollary 4.3.

The sets 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) are trivial when dim(X^)<maxωΘ{|ω|}\dim(\hat{X})<\max_{\omega\in\Theta}\{|\omega|^{\prime}\}.

Proof.

Since codim(𝒫dΘ,𝒫d)=maxωΘ{|ω|}\textup{codim}(\mathcal{P}_{d}^{\Theta},\mathcal{P}_{d})=\max_{\omega\in\Theta}\{|\omega|^{\prime}\} and 𝒫d\mathcal{P}_{d} is contractible, by Theorem 4.2, the claim follows by a general position argument, applied to homotopies of maps from the pair (1+X^(v^),2X^(v^))(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v})) to the pair (𝒫d𝐜Θ,pt)(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt). ∎

Corollary 4.4.

The sets 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}), 𝒬𝒯d,d𝗌𝗎𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{sub}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) are invariants of the path-connected component of the vector field v^\hat{v} in the space 𝖼𝗈𝗇𝗏(X^)\mathsf{conv}(\hat{X}) of convex traversing vector fields.

Proof.

By their definitions, the sets 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) and 𝒬𝒯d,d𝗌𝗎𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{sub}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) depend only of the smooth topological types of the oriented 11-dimensional foliations (v^)\mathcal{L}(\hat{v}), and (v^)\mathcal{L}^{\bullet}(\hat{v}^{\bullet}) on X^\hat{X} and X^×[0,1]\hat{X}\times[0,1], respectively. By Lemma 4.1, the smooth topological types of these foliations do not change along any path in the space 𝖼𝗈𝗇𝗏(X^)\mathsf{conv}(\hat{X}), that contains the point v^\hat{v}. ∎

Corollary 4.5.

The set 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) depends only on the homotopy type of the pair (1+X^(v^),2X^(v^))(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v})).

Proof.

The claim follows instantly from the bijection in formula (4.2). ∎

The next theorem is our main result about the stability of 𝒬𝒯d,d+2𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,\,d+2}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta) as a function of dd in terms of the function ηΘ(d+2)\eta_{\Theta}(d+2) from (2.8).

Theorem 4.3.

(short stabilization: {𝐝𝐝+𝟐}\mathbf{\{d\Rightarrow d+2\}}) Let Θ\Theta be a closed subposet of 𝛀\mathbf{\Omega}. If

dim(X^)<d+2ψΘ(d+2)\dim(\hat{X})<d+2-\psi_{\Theta}(d+2)

and |ω|>2|\omega|^{\prime}>2 for all ωΘd+2]\omega\in\Theta_{\langle d+2]}, then there exists a bijection

Φ𝖾𝗆𝖻:𝒬𝒯d,d+2𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)[(1+X^(v^),2X^(v^)),(𝒫d𝐜Θ,pt)]\displaystyle\Phi^{\mathsf{emb}}:\mathcal{QT}^{\mathsf{emb}}_{d,\,d+2}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta)\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\big{[}(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v})),\,(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt)\big{]}
Proof.

Let k=𝖽𝖾𝖿d+2ψΘ(d+2)k=_{\mathsf{def}}d+2-\psi_{\Theta}(d+2). If |ω|>2|\omega|^{\prime}>2 for all ωΘd+2]\omega\in\Theta_{\langle d+2]}, then both 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta} and 𝒫d+2𝐜Θ\mathcal{P}_{d+2}^{\mathbf{c}\Theta} are simply-connected, since codim(𝒫¯d+2Θ,𝒫¯d+2)>2\textup{codim}(\bar{\mathcal{P}}_{d+2}^{\Theta},\bar{\mathcal{P}}_{d+2})>2 and codim(𝒫¯dΘ,𝒫¯d)>2\textup{codim}(\bar{\mathcal{P}}_{d}^{\Theta},\bar{\mathcal{P}}_{d})>2. By combining Theorem 2.2 with the Alexander duality as in [KSW2], we get an isomorphism ϵ:Hj(𝒫d𝐜Θ;)Hj(𝒫d+2𝐜Θ;)\epsilon_{\ast}:H_{j}(\mathcal{P}_{d}^{\mathbf{c}\Theta};\mathbb{Z})\approx H_{j}(\mathcal{P}_{d+2}^{\mathbf{c}\Theta};\mathbb{Z}) for all j<kj<k. By the Whitehead Theorem (the inverse Hurewicz Theorem) (see Theorem (7.13) in [Wh] or Theorem 10.1 in [Hu]), the map ϵ:𝒫d𝐜Θ𝒫d+2𝐜Θ\epsilon:\mathcal{P}_{d}^{\mathbf{c}\Theta}\to\mathcal{P}_{d+2}^{\mathbf{c}\Theta} is (k1)(k-1)-connected. Thus, if n=dim(1+X^(v^))k1n=\dim(\partial_{1}^{+}\hat{X}(\hat{v}))\leq k-1, then, by the standard application of the obstruction theory, no map Φ:(1+X^(v^),2X^(v^))(𝒫d𝐜Θ,pt)\Phi:(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v}))\to(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt), which is not null-homotopic, becomes null-homotopic in (𝒫d+2𝐜Θ,pt)(\mathcal{P}_{d+2}^{\mathbf{c}\Theta},pt^{\prime}). Now the claim follows from Theorem 4.2. ∎

Theorem 4.3 leads to the following straightforward, but important implication.

Corollary 4.6.

(long stabilization 𝐝\mathbf{d\Rightarrow\infty}) Let Θ\Theta be a closed profinite (see Definition 2.1) poset such that |ω|>2|\omega|^{\prime}>2 for all ωΘ\omega\in\Theta.

Then, given a convex pair (X^,v^)(\hat{X},\hat{v}), the quasiptopy set 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,\,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta) stabilizes towards the set of homotopy classes [(1+X^(v^),2X^(v^)),(𝒫𝐜Θ,pt)]\big{[}(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v})),\,(\mathcal{P}_{\infty}^{\mathbf{c}\Theta},pt)\big{]} for all sufficiently big dd relative to dim(X^)\dim(\hat{X}). \diamondsuit

4.5. Group structure on the ball-shaped convex pseudo-envelops

The connected sum α1α2\alpha_{1}\uplus\alpha_{2} of convex pseudo-envelops α1:X1X^1,α2:X2X^2\alpha_{1}:X_{1}\to\hat{X}_{1},\alpha_{2}:X_{2}\to\hat{X}_{2} can be introduced in a fashion, similar to the operation \uplus in formulae (3.14) and (3.15) from [K9] and formula (3) above.

Let D+n1D^{n-1}_{+} denote the Southern hemisphere in Dn\partial D^{n}. As with the connected sums of submersions from Section 3, there is an ambiguity about how to attach a 11-handle D+n1×[0,1]D^{n-1}_{+}\times[0,1] to 1+X^1(v^1)1+X^2(v^2)\partial_{1}^{+}\hat{X}_{1}(\hat{v}_{1})\coprod\partial_{1}^{+}\hat{X}_{2}(\hat{v}_{2}) and a 11-handle H=Dn×[0,1]H=D^{n}\times[0,1] to X^1X^2\hat{X}_{1}\coprod\hat{X}_{2} to form the “coordinated” connected sums 1+X^1(v^1)#1+X^2(v^2)\partial_{1}^{+}\hat{X}_{1}(\hat{v}_{1})\,\#_{\partial}\,\partial_{1}^{+}\hat{X}_{2}(\hat{v}_{2}) and X^1#X^2\hat{X}_{1}\,\#_{\partial}\,\hat{X}_{2}. The ambiguity arises if 2X^1(v^1)\partial_{2}^{-}\hat{X}_{1}(\hat{v}_{1}) or 2X^2(v^2)\partial_{2}^{-}\hat{X}_{2}(\hat{v}_{2}) has more than a single connected component. To avoid it, as in [K9], formula (3.14), we need to pick a preferred connected component of 2X^1(v^1)\partial_{2}^{-}\hat{X}_{1}(\hat{v}_{1}) and 2X^2(v^2)\partial_{2}^{-}\hat{X}_{2}(\hat{v}_{2}). Using the local models of convex vector fields (as in the proof of Lemma 4.3), the vector fields v^1\hat{v}_{1} and v^2\hat{v}_{2} extend across HH so that the convexity of the extended traversing vector field v^#\hat{v}_{\#_{\partial}} is enforced. For example, the lower diagram in Fig.2 is the connected sum of the upper diagram with itself.

Theorem 4.4.

Let ΘΘ𝛀\Theta^{\prime}\subset\Theta\subset\mathbf{\Omega} be closed subposets which do not contain the element ()(\emptyset). Assume that ddd^{\prime}\geq d and dd0mod2d^{\prime}\equiv d\equiv 0\mod 2. Let v^\hat{v} be a convex traversing vector field on the standard ball Dn+1D^{n+1} such that 1+Dn+1(v^)\partial_{1}^{+}D^{n+1}(\hat{v}) is diffeomorphic to the standard ball DnD^{n}.555The constant vector field will do.

The group operation in the sources of the maps (4.4) and (4.4) below is the connected sum \uplus of convex envelops/convex pseudo-envelops.

\bullet There is a group isomorphism

Φ𝖾𝗆𝖻:𝒬𝒯d,d𝖾𝗆𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)\displaystyle\Phi^{\mathsf{emb}}:\mathcal{QT}^{\mathsf{emb}}_{d,\,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}
(4.7) [[(Dn,Sn1),ϵd,d:(𝒫d𝐜Θ,pt)(𝒫d𝐜Θ,pt)]].\displaystyle\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\big{[}\big{[}(D^{n},\,S^{n-1}),\;\epsilon_{d,\,d^{\prime}}:(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt)\to(\mathcal{P}_{d^{\prime}}^{\mathbf{c}\Theta^{\prime}},pt^{\prime})\big{]}\big{]}.

The group homomorphism

Φ𝗌𝗎𝖻:𝒬𝒯d,d𝗌𝗎𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)\displaystyle\Phi^{\mathsf{sub}}:\mathcal{QT}^{\mathsf{sub}}_{d,\,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\stackrel{{\scriptstyle}}{{\longrightarrow}}
(4.8) epi[[(Dn,Sn1),ϵd,d:(𝒫d𝐜Θ,pt)(𝒫d𝐜Θ,pt)]].\displaystyle\stackrel{{\scriptstyle epi}}{{\longrightarrow}}\big{[}\big{[}(D^{n},\,S^{n-1}),\;\epsilon_{d,\,d^{\prime}}:(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt)\to(\mathcal{P}_{d^{\prime}}^{\mathbf{c}\Theta^{\prime}},pt^{\prime})\big{]}\big{]}.

is an split epimorphism. Moreover, we have a split group extension

(4.9) 1ker(Φ𝗌𝗎𝖻)𝒬𝒯d,d𝗌𝗎𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)𝒬𝒯d,d𝖾𝗆𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)1.\displaystyle\qquad 1\to\ker(\Phi^{\mathsf{sub}})\to\mathcal{QT}^{\mathsf{sub}}_{d,\,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\stackrel{{\scriptstyle\mathcal{R}}}{{\rightarrow}}\mathcal{QT}^{\mathsf{emb}}_{d,\,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\to 1.

\bullet For n>1n>1 all these groups are abelian.

\bullet For n<maxωΘd]{|ω|}n<\max_{\omega\in\Theta_{\langle d]}}\{|\omega|^{\prime}\}, the groups 𝒬𝒯d,d𝖾𝗆𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) are trivial.

Proof.

The main observation is that the construction of the boxes (X^,v^)([0,1]×Y,v~)(\hat{X},\hat{v})\subset([0,1]\times Y,\tilde{v}) in Lemma 4.3 is amenable to the connected sum operation for convex pseudo-envelops of traversing flows. That is, given two boxes ([0,1]×Y1,v~1)(X^1,v^1)([0,1]\times Y_{1},\tilde{v}_{1})\supset(\hat{X}_{1},\hat{v}_{1}) and ([0,1]×Y2,v~2)(X^2,v^2)([0,1]\times Y_{2},\tilde{v}_{2})\supset(\hat{X}_{2},\hat{v}_{2}) as in Lemma 4.3, we get that

(X^1#X^2,v^1#v^2)([0,1]×(Y1#Y2),v~1#v~2)(\hat{X}_{1}\,\#_{\partial}\,\hat{X}_{2},\hat{v}_{1}\,\#_{\partial}\,\hat{v}_{2})\subset([0,1]\times(Y_{1}\,\#_{\partial}\,Y_{2}),\tilde{v}_{1}\,\#_{\partial}\,\tilde{v}_{2})

is also a box as in that lemma. Therefore, all the constructions and arguments from Section 3 in [K9] (like Proposition 3.4, Corollary 3.3, and Theorem 3.2) apply to the convex pseudo-envelops/envelops α:(X,α(v^))(Dn+1,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(D^{n+1},\hat{v}), where the traversing vector field v^\hat{v} is convex with respect to Dn+1\partial D^{n+1}. Thus, as in formula (3) above (see formula (3.15) and Proposition 3.2 from [K9]), for a convex vector field v^\hat{v} with 1+Dn+1(v^)\partial_{1}^{+}D^{n+1}(\hat{v}) being a smooth nn-ball, the quasitopies 𝒬𝒯d,d𝖾𝗆𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) and 𝒬𝒯d,d𝗌𝗎𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{sub}}_{d,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) are groups. For n>1n>1 they are commutative by arguments as in Proposition 3.2 from [K9].

The last claim follows by the general position argument. ∎

Recall that, by Corollary 4.5, the set 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) depends only on the homotopy type of the pair (1+X^(v^),2X^(v^))(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v})); however, the corollary does not make any claims about 𝒬𝒯d,d𝗌𝗎𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{sub}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}). The next proposition, in line with Proposition 4.1 (which deals with the homology nn-spheres, n6n\geq 6), is a hint that 𝒬𝒯d,d𝗌𝗎𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{sub}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) may also depend only on the homotopy type of the pair (1+X^(v^),2X^(v^))(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v})).

Proposition 4.2.

For n5n\geq 5, assuming that the locus 2Dn+1(v^)\partial_{2}^{-}D^{n+1}(\hat{v}) is simply-connected, the groups 𝒬𝒯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{sub/emb}}_{d,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) do not depend on the choice of the convex traversing vector field v^\hat{v} on Dn+1D^{n+1}.

For n3n\leq 3, the groups 𝒬𝒯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{sub/emb}}_{d,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) also do not depend on the convex v^\hat{v}.

Proof.

For a convex v^\hat{v}, the locus 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}) is a deformation retract of Dn+1D^{n+1} and thus has a homotopy type of a point. By the Poincaré duality, 2Dn+1(v^)=(1+X^(v^))\partial_{2}^{-}D^{n+1}(\hat{v})=\partial(\partial_{1}^{+}\hat{X}(\hat{v})) is a homology (n1)(n-1)-sphere Σn1Dn+1\Sigma^{n-1}\subset\partial D^{n+1}. Let us delete a small smooth ball BnB^{n} from the interior of 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}). We denote its complement by WnW^{n}. For n5n\geq 5, assuming that π1(2Dn+1(v^))=1\pi_{1}(\partial_{2}^{-}D^{n+1}(\hat{v}))=1, we may apply the smooth hh-cobordism theorem (see [Mi]) to WW to conclude that it is diffeomorphic to the product Sn1×[0,1]S^{n-1}\times[0,1]. Therefore, 1+Dn+1(v^)\partial_{1}^{+}D^{n+1}(\hat{v}) is a smooth ball DnSnD^{n}\subset S^{n}. Any two regular embeddings DnSnD^{n}\hookrightarrow S^{n} are diffeotopic. By the proof of Lemma 4.1, any diffeotopy of SnS^{n} which maps 1+Dn+1(v^1)\partial_{1}^{+}D^{n+1}(\hat{v}_{1}) to 1+Dn+1(v^2)\partial_{1}^{+}D^{n+1}(\hat{v}_{2}) extends to a diffeotopy of Dn+1D^{n+1} that maps the v^1\hat{v}_{1}-trajectories to v^2\hat{v}_{2}-trajectories, while preserving their orientation. As a result, for n5n\geq 5, the group 𝒬𝒯d,d𝖾𝗆𝖻(Dn+1,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(D^{n+1},\hat{v};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) does not depend on the choice of v^\hat{v}, as long as π1(2Dn+1(v^))=1\pi_{1}(\partial_{2}^{-}D^{n+1}(\hat{v}))=1.

In small dimensions n3n\leq 3, the contractibility of 1+Dn+1(v^)\partial_{1}^{+}D^{n+1}(\hat{v}) implies that it is the standard smooth nn-ball. The case n=1n=1 is obvious. For n=2n=2, if the domain 1+X^(v^)S2\partial_{1}^{+}\hat{X}(\hat{v})\subset S^{2} has a boundary that is a homology 11-sphere, then the domain is the 22-ball, and any 22-balls in S2S^{2} are isotopic. For n=3n=3, if the domain 1+D4(v^)S3\partial_{1}^{+}D^{4}(\hat{v})\subset S^{3} has a smooth boundary that is a homology 22-sphere. By the classification of 22-surfaces, it follows that 2D4(v^)\partial_{2}^{-}D^{4}(\hat{v}) is the standard 22-sphere. By the solution of the 33-dimensional Poincaré Conjecture [P1]-[P3], the contractible domain 1+D4(v^)\partial_{1}^{+}D^{4}(\hat{v}) is with the spherical boundary is the 33-ball, and any two 33-balls in S3S^{3} are isotopic. Thus, by Lemma 4.1, the smooth isotopy type of the locus 1+Dn+1(v^)Sn\partial_{1}^{+}D^{n+1}(\hat{v})\subset S^{n} determines the smooth topological type of the foliation (v^)\mathcal{L}(\hat{v}). Since, any two standard nn-balls 1+X^(v^1)\partial_{1}^{+}\hat{X}(\hat{v}_{1}) and 1+X^(v^2)\partial_{1}^{+}\hat{X}(\hat{v}_{2}) are isotopic in SnS^{n}, the two statements of the proposition are validated. The difficult case n=4n=4 is wide open. ∎

Combining Theorem 4.2 and Theorem 4.4, we get the following claim.

Corollary 4.7.

Let v^\hat{v}^{\parallel} be a constant vector field on the ball Dn+1n+1D^{n+1}\subset\mathbb{R}^{n+1}, and let v^\hat{v} be a convex vector field on X^\hat{X}. Put r=#(π0(2X^(v^)))r=\#(\pi_{0}(\partial_{2}^{-}\hat{X}(\hat{v}))). With the help of the maps from (4.4) and (4.4), the groups

𝒢𝖾𝗆𝖻:=(𝒬𝒯d,d𝖾𝗆𝖻(Dn+1,v^;𝐜Θ;𝐜Θ))r and 𝒢𝗌𝗎𝖻:=(𝒬𝒯d,d𝗌𝗎𝖻(Dn+1,v^;𝐜Θ;𝐜Θ))r\mathcal{G}^{\mathsf{emb}}:=\big{(}\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(D^{n+1},\hat{v}^{\parallel};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\big{)}^{r}\;\text{ and }\;\mathcal{G}^{\mathsf{sub}}:=\big{(}\mathcal{QT}^{\mathsf{sub}}_{d,d^{\prime}}(D^{n+1},\hat{v}^{\parallel};\,\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime})\big{)}^{r}

are represented in the group

(πn(𝒫d𝐜Θ,pt)/ker{(ϵd,d):πn(𝒫d𝐜Θ,pt)πn(𝒫d𝐜Θ,pt)})r.\big{(}\pi_{n}(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt)\big{/}\ker\big{\{}(\epsilon_{d,d^{\prime}})_{\ast}:\pi_{n}(\mathcal{P}_{d}^{\mathbf{c}\Theta},pt)\to\pi_{n}(\mathcal{P}_{d^{\prime}}^{\mathbf{c}\Theta^{\prime}},pt^{\prime})\big{\}}\big{)}^{r}.

We denote by Ψ𝖾𝗆𝖻\Psi^{\mathsf{emb}} and Ψ𝗌𝗎𝖻\Psi^{\mathsf{sub}} these two representations (the first one is an isomorphism).

Then the maps Φ𝗌𝗎𝖻(X^,v^)\Phi^{\mathsf{sub}}(\hat{X},\hat{v}) in (4.2) and Φ𝖾𝗆𝖻(X^,v^)\Phi^{\mathsf{emb}}(\hat{X},\hat{v}) in (4.2) from Theorem 4.2 are equivariant with respect to the 𝒢𝗌𝗎𝖻\mathcal{G}^{\mathsf{sub}}- and 𝒢𝖾𝗆𝖻\mathcal{G}^{\mathsf{emb}}-actions on their source sets and the Ψ𝗌𝗎𝖻(𝒢𝗌𝗎𝖻)\Psi^{\mathsf{sub}}(\mathcal{G}^{\mathsf{sub}})- and Ψ𝖾𝗆𝖻(𝒢𝖾𝗆𝖻)\Psi^{\mathsf{emb}}(\mathcal{G}^{\mathsf{emb}})-actions on their target sets. \diamondsuit

4.6. Quasitopies of envelops with generic combinatorics Θ\Theta and d=dd^{\prime}=d

From now and until Subsection 4.9, each result about quasitopies 𝒬𝒯d,d𝗌𝗎𝖻/𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{sub/emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) of convex envelops is a recognizable immage of a similar result from [K9] about the quasitopies 𝒬𝒯d,d𝗂𝗆𝗆/𝖾𝗆𝖻(Y,Y;𝐜Θ,(),𝐜Θ)\mathcal{QT}^{\mathsf{imm/emb}}_{d,d^{\prime}}(Y,\partial Y;\mathbf{c}\Theta,(\emptyset),\mathbf{c}\Theta^{\prime}) of immersions/embedding into the products ×Y\mathbb{R}\times Y.

Theorem 4.5.

Let (X^,v^)(\hat{X},\hat{v}) be a convex pair. Let d,kd,k be natural numbers such that 2<k<d2<k<d and d0mod2d\equiv 0\mod 2. Put 𝛀||k1=𝖽𝖾𝖿𝐜𝛀||k\mathbf{\Omega}_{|\sim|^{\prime}\leq k-1}=_{\mathsf{def}}\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k}. Then we get a bijection666see (2.12) for the definition of the number A(d,k):=A(d,k,0)A(d,k):=A(d,k,0)

Φ:𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝛀||k1;𝛀||k1)[(1+X^(v^),2X^(v^)),(=1A(d,k)Sk1,)].\displaystyle\Phi:\;\mathcal{QT}^{\mathsf{emb}}_{d,d}(\hat{X},\hat{v};\,\mathbf{\Omega}_{|\sim|^{\prime}\leq k-1};\mathbf{\Omega}_{|\sim|^{\prime}\leq k-1})\;\approx\;\big{[}\big{(}\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v})\big{)},\;\big{(}\bigvee_{\ell=1}^{A(d,\,k)}S^{k-1}_{\ell},\;\star\big{)}\big{]}.
Proof.

The theorem is based on Proposition 3.10 from [K9]. For d>k>2d>k>2, Proposition 2.1, being combined with the Alexander duality, describes the homotopy type of 𝒫d𝛀||k1=𝒫d𝐜𝛀||k\mathcal{P}_{d}^{\mathbf{\Omega}_{|\sim|^{\prime}\leq k-1}}=\mathcal{P}_{d}^{\mathbf{c\Omega}_{|\sim|^{\prime}\geq k}} as a bouquet =1A(d,k)Sk1\bigvee_{\ell=1}^{A(d,\,k)}S^{k-1}_{\ell} of (k1)(k-1)-spheres. The space 𝒫d()\mathcal{P}_{d}^{(\emptyset)} is contractible to the point \star. Thus, by Theorem 4.2 (based on Theorem 3.1 from [K9]), the claim follows. ∎

We consider now the “combinatorially generic” case of convex pseudo-envelops α:(X,α(v^))(X^,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(\hat{X},\hat{v}) for which v^\hat{v} is traversally generic with respect to α(X)\alpha(\partial X). By Definition 4.7, any such α\alpha has tangency patterns that belong to the poset 𝛀||n\mathbf{\Omega}_{|\sim|^{\prime}\leq n}, where n=dimX^1n=\dim\hat{X}-1.

Corollary 4.8.

Let (X^,v^)(\hat{X},\hat{v}) be a convex pair, dimX^=n+1\dim\hat{X}=n+1. For d>n>2d>n>2, we get a bijection

Φ:𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝛀||n;𝛀||n)A(d,n+1).\Phi:\,\mathcal{QT}^{\mathsf{emb}}_{d,d}(\hat{X},\hat{v};\,\mathbf{\Omega}_{|\sim|^{\prime}\leq n};\mathbf{\Omega}_{|\sim|^{\prime}\leq n})\approx\mathbb{Z}^{A(d,\,n+1)}.

These \mathbb{Z}-valued invariants of convex envelops are delivered by the degrees of the maps

{Φ:1+X^(v^)/2X^(v^)Sn}[1,A(d,n+1)],\{\Phi_{\ell}:\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v})\to S^{n}_{\ell}\}_{\ell\in[1,\;A(d,n+1)]},

to the individual spheres, induced by the map Φ\Phi from Theorem 4.5.

In particular, the (d,d;𝛀||n,𝛀||n)(d,d;\mathbf{\Omega}_{|\sim|^{\prime}\leq n},\mathbf{\Omega}_{|\sim|^{\prime}\leq n})-quasitopy class of any traversally generic convex envelop α:(X,α(v^))(X^,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(\hat{X},\hat{v}) is determined by the collection of such degrees.

Proof.

Since X^\hat{X} is connected and v^\hat{v} is convex, the locus 1+X^\partial_{1}^{+}\hat{X} is connected as well.

We repeat the arguments from Proposition 3.6 in [K9]. For n>2n>2, the homotopy classes of the classifying maps Φα:(1+X^(v^),2X^(v^))(𝒫d𝛀||n,𝒫d())\Phi^{\alpha}:(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v}))\to\big{(}\mathcal{P}_{d}^{\mathbf{\Omega}_{|\sim|^{\prime}\leq n}},\,\mathcal{P}_{d}^{(\emptyset)}\big{)} are in 11-to-11 correspondence with homotopy classes of the corresponding maps Φ~α:(1+X^(v^),2X^(v^))(=1A(d,n+1)Sn,)\tilde{\Phi}^{\alpha}:(\partial_{1}^{+}\hat{X}(\hat{v}),\,\partial_{2}^{-}\hat{X}(\hat{v}))\to\big{(}\bigvee_{\ell=1}^{A(d,\,n+1)}S^{n}_{\ell},\,\star\big{)}. For n2n\geq 2, the latter ones are detected by the degrees of the maps {Φα}\{\Phi^{\alpha}_{\ell}\} to the individual spheres {Sn}\{S^{n}_{\ell}\}. ∎

Example 4.5.

Take d=6d=6 and n=3n=3. Then 𝒫¯6𝛀||4\bar{\mathcal{P}}_{6}^{\mathbf{\Omega}_{|\sim|^{\prime}\geq 4}} consist of a single 0-dimensional cell (the ``"𝒫¯6``\infty"\in\bar{\mathcal{P}}_{6}), one 11-dimensional cell, labelled by ω=(6)\omega=(6), and five 22-dimensional cells, labelled by ω=(51),(15),(42),(24),(33)\omega=(51),(15),(42),(24),(33). Hence, A(6,4)=χ(𝒫¯6𝛀||4)1=4A(6,4)=\chi(\bar{\mathcal{P}}_{6}^{\mathbf{\Omega}_{|\sim|^{\prime}\geq 4}})-1=4. The space 𝒫¯6𝛀||4\bar{\mathcal{P}}_{6}^{\mathbf{\Omega}_{|\sim|^{\prime}\geq 4}} has a homotopy type of a bouquet =14S2\bigvee_{\ell=1}^{4}S^{2} of four 22-spheres. By the Alexander duality, 𝒫6𝐜𝛀||4=𝒫6𝛀||3\mathcal{P}_{6}^{\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq 4}}=\mathcal{P}_{6}^{\mathbf{\Omega}_{|\sim|^{\prime}\leq 3}} has a homotopy type of a bouquet =14S3\bigvee_{\ell=1}^{4}S^{3} of four 33-spheres.

We pick a convex and traversing vector field v^\hat{v} on D4D^{4}. Recall that 1+D4(v^)\partial_{1}^{+}D^{4}(\hat{v}) is contractible and thus 2D4(v^)\partial_{2}^{-}D^{4}(\hat{v}) a homology 22-sphere, which implies that 2D4(v^)\partial_{2}^{-}D^{4}(\hat{v}) is diffeomorphic to S2S3S^{2}\subset S^{3}. Thus 1+D4(v^)\partial_{1}^{+}D^{4}(\hat{v}), by [P1], [P2], is diffeomorphic to the ball D3D^{3}. Therefore, for any convex v^\hat{v}, we get the group isomorphism

𝒬𝒯6,6𝖾𝗆𝖻(D4,v^;𝛀||3;𝛀||3)π3(=14S3)4.\mathcal{QT}^{\mathsf{emb}}_{6,6}(D^{4},\hat{v};\,\mathbf{\Omega}_{|\sim|^{\prime}\leq 3};\mathbf{\Omega}_{|\sim|^{\prime}\leq 3})\approx\pi_{3}(\bigvee_{\ell=1}^{4}S^{3}_{\ell})\approx\mathbb{Z}^{4}.\quad

As a result, any element [α]𝒬𝒯6,6𝗌𝗎𝖻(D4,v^;𝛀||3;𝛀||3)[\alpha]\in\mathcal{QT}^{\mathsf{sub}}_{6,6}(D^{4},\hat{v};\,\mathbf{\Omega}_{|\sim|^{\prime}\leq 3};\mathbf{\Omega}_{|\sim|^{\prime}\leq 3}) generates four integer-valued characteristic invariants. For embeddings α\alpha, they determine [α][\alpha].

At the same time, by Corollary 4.9 below, 𝒬𝒯6,8𝖾𝗆𝖻(D4,v^;𝛀||3;𝛀||4)=0\mathcal{QT}^{\mathsf{emb}}_{6,8}(D^{4},\hat{v};\,\mathbf{\Omega}_{|\sim|^{\prime}\leq 3};\mathbf{\Omega}_{|\sim|^{\prime}\leq 4})=0. \diamondsuit

The following claim contrasts Corollary 4.8.

Corollary 4.9.

Under the hypotheses of Corollary 4.8, including d>n>2d>n>2, the set 𝒬𝒯d,d+2𝖾𝗆𝖻(X^,v^;𝛀||n;𝛀||n+1)\mathcal{QT}^{\mathsf{emb}}_{d,\,d+2}(\hat{X},\hat{v};\,\mathbf{\Omega}_{|\sim|^{\prime}\leq n};\mathbf{\Omega}_{|\sim|^{\prime}\leq n+1}) consists of a single element, represented by X=X=\emptyset.

In particular, the (d,d+2;𝛀||n,𝛀||n+1)(d,d+2;\mathbf{\Omega}_{|\sim|^{\prime}\leq n},\mathbf{\Omega}_{|\sim|^{\prime}\leq n+1})-quasitopy class of any traversally generic convex envelop α:(X,α(v^))(X^,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(\hat{X},\hat{v}) is trivial.

Proof.

Consider a convex pseudo-envelop α:(X,α(v^))(X^,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(\hat{X},\hat{v}) as in Theorem 4.5.

The claim is based on the observation that any map from the nn-dimensional CWCW-complex 1+X^(v^)/2X^(v^)\partial_{1}^{+}\hat{X}(\hat{v})/\partial_{2}^{-}\hat{X}(\hat{v}) to 𝒫d𝐜𝛀||n\mathcal{P}_{d}^{\mathbf{c\Omega}_{|\sim|^{\prime}\geq n}}—homotopically a bouquet of nn-spheres— is null-homptopic in 𝒫d+2𝐜𝛀||n+1\mathcal{P}_{d+2}^{\mathbf{c\Omega}_{|\sim|^{\prime}\geq n+1}}, since the latter space has the homotopy type of a bouquet of (n+1)(n+1)-spheres. Again, by Theorem 4.2, the pseudo-envelop α\alpha is null-quasitopic. ∎

4.7. From inner framed cobordisms of 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}) to quasitopies of kk-flat convex envelops

Theorem 4.5 suggests a somewhat unexpected connection between quasitopies of certain convex envelops α:XX^\alpha:X\to\hat{X} and inner framed cobordisms of the manifold X^\partial\hat{X}.

Let YY be a compact smooth nn-manifold. For k1nk-1\leq n, we associate with YY the set of inner framed cobordisms Ξk1(Y)\mathcal{FB}^{k-1}_{\Xi}(Y). These cobordisms are based on codimension k1k-1 smooth closed submanifolds ZZ of YY of the form Z=σ=1A(d,k,q)ZσZ=\coprod_{\sigma=1}^{A(d,k,q)}Z_{\sigma}, where the number A(d,k,q)A(d,k,q) is introduced in (2.12). The normal (k1)(k-1)-bundle ν(Z,Y)\nu(Z,Y) is required to be framed. The disjoint “components” {Zσ}\{Z_{\sigma}\} of ZZ are marked with different colors σ\sigma from a pallet Ξ\Xi of cardinality A(d,k,q)A(d,k,q).

We have seen in [K9], Proposition 3.11, that the inner framed Ξ\Xi-colored codimension k1k-1 bordisms of the nn-dimensional manifold YY produce, via the Thom construction, quite special kk-flat embeddings β:M×Y\beta:M\to\mathbb{R}\times Y, where dimM=dimY\dim M=\dim Y. The analogous mechanism, with the help of Lemma 4.3 and Theorem 4.5, generates special kk-flat envelops (see Definition 4.6).

Proposition 4.3.

Let X^\hat{X} be a smooth compact connected (n+1)(n+1)-dimensional manifold, equipped with a convex traversing vector field v^\hat{v}. Let k[3,n+1]k\in[3,n+1], k<dk<d, and d0mod2d\equiv 0\mod 2. Then the Thom construction delivers a bijection

Th:Ξk1(1+X^(v^))𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜𝛀||k;𝐜𝛀||k),Th:\mathcal{FB}_{\Xi}^{k-1}(\partial_{1}^{+}\hat{X}(\hat{v}))\approx\mathcal{QT}^{\mathsf{emb}}_{d,d}(\hat{X},\hat{v};\,\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k};\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k}),

where Ξk1(1+X^(v^))\mathcal{FB}_{\Xi}^{k-1}(\partial_{1}^{+}\hat{X}(\hat{v})) denotes the set of inner framed Ξ\Xi-colored (nk+1)(n-k+1)-dimensional bordisms of the space 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}). Here #Ξ=A(d,k)\#\Xi=A(d,k) (see 2.12). \diamondsuit

Example 4.6.

Let us recycle Example 3.5 from [K9]: take k=3k=3, d=6d=6, q=0q=0, and n=3n=3. Then #Ξ=10\#\Xi=10. So we get a homotopy equivalence τ:𝒫6𝐜Θ||3σ=110Sσ2\tau:\mathcal{P}_{6}^{\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3}}\sim\bigvee_{\sigma=1}^{10}S^{2}_{\sigma}. Let (X^,v^)(\hat{X},\hat{v}) be a convex pair, where dimX^=4\dim\hat{X}=4, and let Z1+X^(v^)Z\subset\partial_{1}^{+}\hat{X}(\hat{v}) be a normally framed 11-dimensional closed submanifold, each loop in ZZ being colored with a color from the pallet Ξ\Xi. Using the isomorphism

Th:Ξ2(1+X^(v^))𝖰𝖳6,6𝖾𝗆𝖻(X^,v^;𝐜Θ||3;𝐜Θ||3),Th:\mathcal{FB}^{2}_{\Xi}(\partial_{1}^{+}\hat{X}(\hat{v}))\approx\mathsf{QT}^{\mathsf{emb}}_{6,6}\big{(}\hat{X},\hat{v};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3}\big{)},

any framed link Z1+X^(v^)Z\subset\partial_{1}^{+}\hat{X}(\hat{v}), colored with 1010 distinct colors at most, produces a quasitopy class of a convex envelop α:(X,α(v^))(X^,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(\hat{X},\hat{v}). Its tangency patterns ω\omega, except for ω=()\omega=(\emptyset), have only entries from the list {1,2,3}\{1,2,3\}, so that no more than one 33 is present in ω\omega, and no more than two 22’s are present, while |ω|6|\omega|\leq 6.

Now, let us assume that D4D^{4} carries a constant vector field vv^{\parallel} and that 2X^(v^)\partial_{2}^{-}\hat{X}(\hat{v})\neq\emptyset. Note that any element of H1(1+X^(v^));)H_{1}(\partial_{1}^{+}\hat{X}(\hat{v}));\mathbb{Z}) my be realized by a disjoint union of framed oriented loops. Then, like in Example 3.5 from [K9], the orbit-space of the 𝖦6,6𝖾𝗆𝖻(D4;𝐜Θ||3;𝐜Θ||3)\mathsf{G}^{\mathsf{emb}}_{6,6}(D^{4};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3})-action on 𝖰𝖳6,6𝖾𝗆𝖻(X^,v^;𝐜Θ||3;𝐜Θ||3)\mathsf{QT}^{\mathsf{emb}}_{6,6}\big{(}\hat{X},\hat{v};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3}\big{)} admits a surjection on the group (H1(1+X^(v^));))10(H1(X^;))10(H_{1}(\partial_{1}^{+}\hat{X}(\hat{v}));\mathbb{Z}))^{10}\approx(H_{1}(\hat{X};\mathbb{Z}))^{10}, provided that X^\partial\hat{X} is orientable.

Let us give a couple of specific examples of this fact. Consider the box T3×[0,1]T^{3}_{\circ}\times[0,1], where T3T^{3}_{\circ} is the compliment in the 33-torus T3T^{3} to a ball D3D^{3}. By rounding the corners of the box, we get a convex pair (X^,v^)(\hat{X},\hat{v}), where X^\hat{X} is homeomorpfic to T3×[0,1]T^{3}_{\circ}\times[0,1]. Then the orbit-space of the 𝖦6,6𝖾𝗆𝖻(D4;𝐜Θ||3;𝐜Θ||3)\mathsf{G}^{\mathsf{emb}}_{6,6}(D^{4};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3})-action on 𝖰𝖳6,6𝖾𝗆𝖻(X^,v^;𝐜Θ||3;𝐜Θ||3)\mathsf{QT}^{\mathsf{emb}}_{6,6}\big{(}\hat{X},\hat{v};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3}\big{)} is mapped onto the lattice 30\mathbb{Z}^{30}.

Let M=3/ΓM=\mathbb{H}^{3}/\Gamma be a compact hyperbolic 33-manifold and M:=MD3M_{\circ}:=M\setminus D^{3}. By rounding the corners of the box M×[0,1]M_{\circ}\times[0,1], we get a convex pair (X^,v^)(\hat{X},\hat{v}). Then the orbit-space of the 𝖦6,6𝖾𝗆𝖻(D4;𝐜Θ||3;𝐜Θ||3)\mathsf{G}^{\mathsf{emb}}_{6,6}(D^{4};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3})-action on 𝖰𝖳6,6𝖾𝗆𝖻(X^,v^;𝐜Θ||3;𝐜Θ||3)\mathsf{QT}^{\mathsf{emb}}_{6,6}\big{(}\hat{X},\hat{v};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3};\mathbf{c}\Theta_{|\sim|^{\prime}\geq 3}\big{)} is mapped onto the abelian group (Γ/[Γ,Γ])10(\Gamma/[\Gamma,\Gamma])^{10}, where [Γ,Γ][\Gamma,\Gamma] denotes the commutator. \diamondsuit

By Proposition 4.3 and repeating the (obviously modified) arguments in the proof of Corollary 3.13 from [K9], we get the following corollary.

Corollary 4.10.

Let d>kd>k, and d0mod2d\equiv 0\mod 2. Let (X^,v^)(\hat{X},\hat{v}) be a convex pair.

  • For dimX^=k3\dim\hat{X}=k\geq 3 and any choice of κπ0(2X^(v^))\kappa\in\pi_{0}(\partial_{2}^{-}\hat{X}(\hat{v})), the group 𝒬𝒯d,d𝖾𝗆𝖻(Dk,v^;𝐜𝛀||k;𝐜𝛀||k)\mathcal{QT}^{\mathsf{emb}}_{d,d}(D^{k},\hat{v}^{\parallel};\,\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k};\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k}) acts freely and transitively on the set 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜𝛀||k;𝐜𝛀||k)\mathcal{QT}^{\mathsf{emb}}_{d,d}(\hat{X},\hat{v};\,\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k};\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k}). Thus both sets are in a 11-to-11 correspondence.

  • For dimX^=k+1>5\dim\hat{X}=k+1>5, a simply-connected X^\hat{X}, and any choice of κπ0(2X^(v^))\kappa\in\pi_{0}(\partial_{2}^{-}\hat{X}(\hat{v})), the group 𝒬𝒯d,d𝖾𝗆𝖻(Dk+1,v^;𝐜𝛀||k;𝐜𝛀||k)\mathcal{QT}^{\mathsf{emb}}_{d,d}(D^{k+1},\hat{v}^{\parallel};\,\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k};\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k}) acts freely and transitively on the set 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜𝛀||k;𝐜𝛀||k)\mathcal{QT}^{\mathsf{emb}}_{d,d}(\hat{X},\hat{v};\,\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k};\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq k}). Again, both sets are in a 11-to-11 correspondence.

    \diamondsuit

Example 4.7.

Let X^=2×D2\hat{X}=\mathbb{C}\mathbb{P}^{2}\times D^{2} with v^\hat{v} on 2×D2\mathbb{C}\mathbb{P}^{2}\times D^{2} being generated by a constant vector field on D2D^{2}. Then 𝒬𝒯d,d𝖾𝗆𝖻(D6,v^;𝐜𝛀||5;𝐜𝛀||5)\mathcal{QT}^{\mathsf{emb}}_{d,d}(D^{6},\hat{v}^{\parallel};\,\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq 5};\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq 5}) acts freely and transitively on the set 𝒬𝒯d,d𝖾𝗆𝖻(2×D2,v^;𝐜𝛀||5;𝐜𝛀||5)\mathcal{QT}^{\mathsf{emb}}_{d,d}(\mathbb{C}\mathbb{P}^{2}\times D^{2},\hat{v};\,\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq 5};\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq 5}) for all even d6d\geq 6. Thus both sets are in 11-to-11 correspondence. As a result, 𝒬𝒯d,d𝖾𝗆𝖻(2×D2,v^;𝐜𝛀||5;𝐜𝛀||5)\mathcal{QT}^{\mathsf{emb}}_{d,d}(\mathbb{C}\mathbb{P}^{2}\times D^{2},\hat{v};\,\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq 5};\mathbf{c}\mathbf{\Omega}_{|\sim|^{\prime}\geq 5}) acquires the structure of the abelian group π5(i=1A(d,5)Si4,)\pi_{5}\big{(}\bigvee_{i=1}^{A(d,5)}S^{4}_{i},\,\star\big{)}. \diamondsuit

4.8. Convex envelops with special combinatorics Θ\Theta

For a given ω𝛀\omega\in\mathbf{\Omega}, we denote by ω\langle\omega\rangle the minimal closed poset, generated by ω\omega. Combining Theorem 4.2 and Theorem 3.5 from [K9] with Proposition 4.2, we get the following result, in which the constraint d12d\leq 12 reflects only the scope of the numerical experiments in [KSW2].

Theorem 4.6.

Assume that d12d\leq 12 and d0mod2d\equiv 0\mod 2. Let ω𝛀d\omega\in\mathbf{\Omega}_{\leq d} be such that |ω|>2|\omega|^{\prime}>2. Let X^\hat{X} be a smooth compact connected (n+1)(n+1)-dimensional manifold, equipped with a convex traversing vector field v^\hat{v}.

Then the quasitopy set 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜ω;𝐜ω)\mathcal{QT}^{\mathsf{emb}}_{d,d}(\hat{X},\hat{v};\mathbf{c}\langle\omega\rangle;\mathbf{c}\langle\omega\rangle) either consists of single element (is trivial), or is isomorphic to the cohomotopy set πk(1+X^(v^)/2X^(v^))\pi^{k}(\partial_{1}^{+}\hat{X}(\hat{v})/\partial_{2}^{-}\hat{X}(\hat{v})), where k=k(ω)[|ω|1,d1]k=k(\omega)\in[|\omega|^{\prime}-1,\,d-1].

In particular, for a constant vector field v^=v\hat{v}=v^{\parallel}, d12d\leq 12, and ω\omega and k=k(ω)k=k(\omega) from the table in [K9], Appendix, the group 𝒬𝒯d,d𝖾𝗆𝖻(Dn+1,v^;𝐜ω,𝐜ω)πn(Sk)\mathcal{QT}^{\mathsf{emb}}_{d,d}(D^{n+1},\hat{v};\mathbf{c}\langle\omega\rangle,\mathbf{c}\langle\omega\rangle)\approx\pi_{n}(S^{k}). A similar claim is valid for any convex v^\hat{v} on Dn+1D^{n+1}, provided that either n3n\leq 3 or n5n\geq 5 and 2X^(v^)\partial_{2}^{-}\hat{X}(\hat{v}) is simply-connected. \diamondsuit

For d13d\leq 13, the table from Appendix in [K9] lists all ω\omega’s and the corresponding k=k(ω)k=k(\omega)’s for which 𝒫d𝐜ω\mathcal{P}_{d}^{\mathbf{c}\langle\omega\rangle} is homologically nontrivial. In fact, 𝒫d𝐜ω\mathcal{P}_{d}^{\mathbf{c}\langle\omega\rangle} is a homology kk-sphere and even a homotopy kk-sphere, at least when |ω|>2|\omega|^{\prime}>2 [K9], a quite mysterious phenomenon…

Let us recycle Example 3.4 from [K9], while adapting it to convex envelops.

Example 4.8.

Let d=8d=8 and ω=(4)\omega=(4). Then, for any convex pair (X^,v^)(\hat{X},\hat{v}), using the list in [K9], Appendix, we get a bijection

𝒬𝒯8,8𝖾𝗆𝖻(X^,v^;𝐜(4),𝐜(4))π4(1+X^(v^)/2X^(v^)),\mathcal{QT}^{\mathsf{emb}}_{8,8}(\hat{X},\hat{v};\mathbf{c}\langle(4)\rangle,\mathbf{c}\langle(4)\rangle)\approx\pi^{4}\big{(}\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v})\big{)},

where π4()\pi^{4}(\sim) stands for the 44-cohomotopy set. In particular, using Proposition 4.2, for any convex v^\hat{v} on D8D^{8} such that the locus 2D8(v^)\partial_{2}^{-}D^{8}(\hat{v}) is simply-connected (the constant v^\hat{v} will do), we get a group isomorphism:

𝒬𝒯8,8𝖾𝗆𝖻(D8,v^;𝐜(4),𝐜(4))π7(S4)×12.\mathcal{QT}^{\mathsf{emb}}_{8,8}(D^{8},\hat{v};\mathbf{c}\langle(4)\rangle,\mathbf{c}\langle(4)\rangle)\approx\pi_{7}(S^{4})\approx\mathbb{Z}\times\mathbb{Z}_{12}.

Let d=12d=12 and ω=(11213)\omega=(11213). Then, for any for any convex pair (X^,v^)(\hat{X},\hat{v}), by the same list from [K9], we get a bijection

𝒬𝒯12,12𝖾𝗆𝖻(X^,v^;𝐜(11213),𝐜(11213))π6(1+X^(v^)/2X^(v^)).\mathcal{QT}^{\mathsf{emb}}_{12,12}(\hat{X},\hat{v};\mathbf{c}\langle(11213)\rangle,\mathbf{c}\langle(11213)\rangle)\approx\pi^{6}\big{(}\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v})\big{)}.

Again, using Proposition 4.2, for any convex v^\hat{v} on D10D^{10}, such that the locus 2D10(v^)\partial_{2}^{-}D^{10}(\hat{v}) is simply-connected (again, the constant v^\hat{v} will do), we get a group isomorphism

𝖰𝖳12,12𝖾𝗆𝖻(D10,v^;𝐜(11213),𝐜(11213))π9(S6)24.\mathsf{QT}^{\mathsf{emb}}_{12,12}(D^{10},\hat{v};\mathbf{c}\langle(11213)\rangle,\mathbf{c}\langle(11213)\rangle)\approx\pi_{9}(S^{6})\approx\mathbb{Z}_{24}.\quad\diamondsuit

Proposition 4.4 below deals with special Θ\Theta’s for which 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta} are K(π,1)K(\pi,1)-spaces and π\pi is a free group.

Let H,GH,G be two groups, and 𝖧𝗈𝗆(H,G)\mathsf{Hom}(H,G) be the group of their homomorphisms. Then GG acts on 𝖧𝗈𝗆(H,G)\mathsf{Hom}(H,G) by the conjugation: for any homomorphism ϕ:HG\phi:H\to G, hHh\in H, and gGg\in G, we define (Adgϕ)(h)(Ad_{g}\phi)(h) by the formula g1ϕ(h)gg^{-1}\phi(h)g. We denote by 𝖧𝗈𝗆(H,G)\mathsf{Hom}^{\bullet}(H,G) the quotient 𝖧𝗈𝗆(H,G)/AdG\mathsf{Hom}(H,G)/Ad_{G}.

Proposition 4.4.

Let 𝐜Θ\mathbf{c}\Theta consist of all ω\omega’s with entries 11 and 22 only and no more than a single entry 22. Put κ(d)=𝖽𝖾𝖿d(d2)4\kappa(d)=_{\mathsf{def}}\frac{d(d-2)}{4} for d0mod2d\equiv 0\mod 2. We denote by 𝖥κ(d)\mathsf{F}_{\kappa(d)} the free group in κ(d)\kappa(d) generators.

Consider a convex pair (X^,v^)(\hat{X},\hat{v}), where X^\hat{X} is connected. If 2X^(v^)\partial_{2}^{-}\hat{X}(\hat{v})\neq\emptyset, then there is a bijection

Φ𝖾𝗆𝖻:𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)𝖧𝗈𝗆(π1(X^),𝖥κ(d))\Phi^{\mathsf{emb}}:\mathcal{QT}^{\mathsf{emb}}_{d,d}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta)\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\mathsf{Hom}(\pi_{1}(\hat{X}),\mathsf{F}_{\kappa(d)})

and a surjection

Φ𝗌𝗎𝖻:𝒬𝒯d,d𝗌𝗎𝖻(X^,v^;𝐜Θ;𝐜Θ)𝖧𝗈𝗆(π1(X^),𝖥κ(d)).\Phi^{\mathsf{sub}}:\mathcal{QT}^{\mathsf{sub}}_{d,d}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta)\longrightarrow\mathsf{Hom}(\pi_{1}(\hat{X}),\mathsf{F}_{\kappa(d)}).

When 2X^(v^)=\partial_{2}^{-}\hat{X}(\hat{v})=\emptyset, then similar claims hold with the target of Φ𝖾𝗆𝖻\Phi^{\mathsf{emb}} and Φ𝗌𝗎𝖻\Phi^{\mathsf{sub}} being replaced by the set 𝖧𝗈𝗆(π1(X^),𝖥κ(d))\mathsf{Hom}^{\bullet}(\pi_{1}(\hat{X}),\mathsf{F}_{\kappa(d)}).

In particular, Φ𝖾𝗆𝖻:𝒬𝒯d,d𝖾𝗆𝖻(S1×[0,1],v^;𝐜Θ,𝐜Θ)𝖥κ(d)/Ad𝖥κ(d)\Phi^{\mathsf{emb}}:\mathcal{QT}^{\mathsf{emb}}_{d,d}(S^{1}\times[0,1],\hat{v};\mathbf{c}\Theta,\mathbf{c}\Theta)\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\mathsf{F}_{\kappa(d)}/Ad_{\mathsf{F}_{\kappa(d)}}, the free group of cyclic words in κ(d)\kappa(d) letters (see Fig.3 from [K9]). Here v^\hat{v} is tangent to the fibers of the obvious projection S1×[0,1]S1.S^{1}\times[0,1]\to S^{1}.

If π1(X^)\pi_{1}(\hat{X}) has no free images, then the group 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ,𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d}(\hat{X},\hat{v};\mathbf{c}\Theta,\mathbf{c}\Theta) is trivial.

Proof.

We observe that π1(1+X^(v^)π1(X^)\pi_{1}(\partial_{1}^{+}\hat{X}(\hat{v})\approx\pi_{1}(\hat{X}) since the two spaces are homotopy equivalent due to the convexity of v^\hat{v}. By Lemma 4.3, we can incapsulate (X^,v^)(\hat{X},\hat{v}) in a box ×Y\mathbb{R}\times Y, where YY is homeomorphic to 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}). Using this YY, the claim follows from Corollary 3.12 in [K9]. ∎

The next proposition is a stabilization result, by the increasing ddd^{\prime}\geq d, for the convex envelops with (using the terminology of [Ar]), kk-moderate tangent patterns ω𝐜Θmaxk\omega\in\mathbf{c}\Theta_{\max\geq k}. The entries of such ω\omega’s are all less than kk. Proposition 4.5 below follows instantly from [K9], Proposition 3.8, by combining it with Lemma 4.3.

Proposition 4.5.

Let k4k\geq 4. If dimY(k2)(d/k+1)2\dim Y\leq(k-2)(\lceil d/k\rceil+1)-2, then the classifying map

Φ𝖾𝗆𝖻:𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θmaxk;𝐜Θmaxk)[(1+X^(v^),2X^(v^)),(𝒫d𝐜Θmaxk,pt)]\displaystyle\qquad\qquad\Phi^{\mathsf{emb}}:\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta_{\max\geq k};\mathbf{c}\Theta_{\max\geq k})\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\big{[}(\partial_{1}^{+}\hat{X}(\hat{v}),\partial_{2}^{-}\hat{X}(\hat{v})),(\mathcal{P}_{d}^{\mathbf{c}\Theta_{\max\geq k}},pt)\big{]}

is a bijection, and the classifying map

Φ𝗌𝗎𝖻:𝒬𝒯d,d𝗂𝗆𝗆(X^,v^;𝐜Θmaxk;𝐜Θmaxk)[(1+X^(v^),2X^(v^)),(𝒫d𝐜Θmaxk,pt)]\displaystyle\qquad\qquad\Phi^{\mathsf{sub}}:\mathcal{QT}^{\mathsf{imm}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta_{\max\geq k};\mathbf{c}\Theta_{\max\geq k})\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\big{[}(\partial_{1}^{+}\hat{X}(\hat{v}),\partial_{2}^{-}\hat{X}(\hat{v})),(\mathcal{P}_{d}^{\mathbf{c}\Theta_{\max\geq k}},pt)\big{]}

is a surjection for any ddd^{\prime}\geq d, ddmod2d^{\prime}\equiv d\mod 2.

In particular, for a given (X^,v^)(\hat{X},\hat{v}), the quasitopy 𝖰𝖳d,d𝖾𝗆𝖻(X^,v^;𝐜Θmaxk;𝐜Θmaxk)\mathsf{QT}^{\mathsf{emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta_{\max\geq k};\mathbf{c}\Theta_{\max\geq k}) stabilizes for all ddkk2(dimX^+3k)d^{\prime}\geq d\geq\frac{k}{k-2}(\dim\hat{X}+3-k), a linear function in dimX^\dim\hat{X}. \diamondsuit

Example 4.9.

Let k=4k=4. By Proposition 4.5, for any compact connected 33-dimensional convex pair (X^,v^)(\hat{X},\hat{v}), we get bijections:

𝒬𝒯4,4𝖾𝗆𝖻(X^,v^;𝐜Θmax4;𝐜Θmax4)π2(1+X^(v^)/2X^(v^)),\displaystyle\mathcal{QT}^{\mathsf{emb}}_{4,4}(\hat{X},\hat{v};\mathbf{c}\Theta_{\max\geq 4};\mathbf{c}\Theta_{\max\geq 4})\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\pi^{2}(\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v})),
𝒬𝒯4,6𝖾𝗆𝖻(X^,v^;𝐜Θmax4;𝐜Θmax4)π2(1+X^(v^)/2X^(v^)),\displaystyle\mathcal{QT}^{\mathsf{emb}}_{4,6}(\hat{X},\hat{v};\mathbf{c}\Theta_{\max\geq 4};\mathbf{c}\Theta_{\max\geq 4})\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\pi^{2}(\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v})),
𝒬𝒯6,6𝖾𝗆𝖻(X^,v^;𝐜Θmax4;𝐜Θmax4)π2(1+X^(v^)/2X^(v^)),\displaystyle\mathcal{QT}^{\mathsf{emb}}_{6,6}(\hat{X},\hat{v};\mathbf{c}\Theta_{\max\geq 4};\mathbf{c}\Theta_{\max\geq 4})\stackrel{{\scriptstyle\approx}}{{\longrightarrow}}\pi^{2}(\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v})),

whose target is the second cohomotopy group π2(1+X^(v^)/2X^(v^))\pi^{2}(\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v})) of the singular connected surface 1+X^(v^)/2X^(v^)\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v}). This cohomotopy group is isomorphic to \mathbb{Z} via the degree invariant. Thus, all the three types of quasitopy classes are determined by this degree. Assuming 2X^(v^)=\partial_{2}^{-}\hat{X}(\hat{v})=\emptyset, we may replace π2(1+X^(v^)/2X^(v^))\pi^{2}(\partial_{1}^{+}\hat{X}(\hat{v})\big{/}\partial_{2}^{-}\hat{X}(\hat{v})) by π2(X^)\pi^{2}(\hat{X}). \diamondsuit

4.9. Characteristic classes of convex pseudo-evelops

In this subsection, we will use the cohomology H(𝒫d𝐜Θ)H^{\ast}(\mathcal{P}_{d}^{\mathbf{c}\Theta}) of the classifying space 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta} (see [KSW2] and Section 2) to produce a variety of characteristic classes of convex pseudo-envelops.

Since any convex pseudo-envelop α:(X,α(v^))(X^,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(\hat{X},\hat{v}) produces an immersion α:XX^\alpha^{\partial}:\partial X\to\hat{X}, the following theorem follows directly from Theorem 3.3, [K9].

Theorem 4.7.

Let (X^,v^)(\hat{X},\hat{v}) be a convex pair. Pick d=dd^{\prime}=d, d0mod2d\equiv 0\mod 2, and Θ=Θ\Theta^{\prime}=\Theta.

Then any convex pseudo-envelop α:(X,α(v^))(X^,v^)\alpha:(X,\alpha^{\dagger}(\hat{v}))\to(\hat{X},\hat{v}) induces a characteristic homomorphism (Φα)(\Phi^{\alpha})^{\ast} from the \ast-homology of the differential complex

{(#):[Θd]#][Θd]#]},\big{\{}(\partial^{\#})^{\ast}:\mathbb{Z}[\Theta_{\langle d]}^{\#}]^{\ast}\to\mathbb{Z}[\Theta_{\langle d]}^{\#}]^{\ast}\big{\}},

dual to the differential complex in (2.5), to the cohomology H(X^;)H^{\ast}(\hat{X};\mathbb{Z}), and, via α\alpha^{\ast}, further to the cohomology H(X;)H^{\ast}(X;\mathbb{Z}).

The (d,d;𝐜Θ,𝐜Θ)(d,d;\mathbf{c}\Theta,\mathbf{c}\Theta)-quasitopic pseudo-envelops/envelops induce the same characteristic homomorphisms. \diamondsuit

Let us revisit the Arnold-Vassiliev case [Ar], [V] of real polynomials with moderate singularities. Let Θmaxk𝛀d]\Theta_{\max\,\geq k}\subset\mathbf{\Omega}_{\langle d]} be the closed poset, consisting of ω\omega’s with the maximal entry k\geq k. For k3k\geq 3, the cohomology Hj(𝒫𝐜Θmaxk,pt;)H^{j}(\mathcal{P}^{\mathbf{c}\Theta_{\max\geq k}},pt;\mathbb{Z}) is isomorphic to \mathbb{Z} in each dimension jj of the form (k2)m(k-2)m, where the integer md/km\leq d/k, and vanishes otherwise [Ar]. The cohomology ring H(𝒫𝐜Θmaxk,pt;)H^{\ast}(\mathcal{P}^{\mathbf{c}\Theta_{\max\geq k}},pt;\mathbb{Z}) was computed by Vassiliev in [V], Theorem 1 on page 87. Here is the summery of his result: consider the graded ring 𝒱assd,k\mathcal{V}ass_{d,k}, multiplicatively generated over \mathbb{Z} by the elements {em}md/k\{e_{m}\}_{m\leq d/k} of the degrees deg(em)=m(k2)\deg(e_{m})=m(k-2), subject to the relations

(4.10) elem=(l+m)!l!m!el+m for k0mod2, and the relations\displaystyle e_{l}\cdot e_{m}=\frac{(l+m)!}{l!\cdot m!}\;e_{l+m}\text{\; for }k\equiv 0\mod 2,\text{ and the relations }
e1e1=0,e1e2m=e2m+1,\displaystyle e_{1}\cdot e_{1}=0,\quad e_{1}\cdot e_{2m}=e_{2m+1},\
(4.11) e2le2m=(l+m)!l!m!e2l+2m for k1mod2.\displaystyle e_{2l}\cdot e_{2m}=\frac{(l+m)!}{l!\cdot m!}\;e_{2l+2m}\text{\; for }k\equiv 1\mod 2.

Combining Lemma 4.3 with Proposition 3.7 from [K9], we get the following assertion.

Proposition 4.6.

Let k3k\geq 3. Consider a convex pseudo-envelop α:(X,v)(X^,v^)\alpha:(X,v)\to(\hat{X},\hat{v}) whose tangency patterns to the v^\hat{v}-flow belong to 𝐜Θmaxk𝛀d]\mathbf{c}\Theta_{\max\,\geq k}\subset\mathbf{\Omega}_{\langle d]} (the tangencies are kk-moderate).

Then α\alpha generates a characteristic ring homomorphism

(Φα):𝒱assd,kH(1+X^(v^),2X^(v^);),(\Phi^{\alpha})^{\ast}:\mathcal{V}ass_{d,k}\to H^{\ast}(\partial_{1}^{+}\hat{X}(\hat{v}),\partial_{2}^{-}\hat{X}(\hat{v});\mathbb{Z}),

which is an invariant of the quasitopy class of α\alpha. In other words, we get a map

Φd,k:𝒬𝒯d,d𝗂𝗆𝗆(X^,v^;𝐜Θmaxk;𝐜Θmaxk)𝖧𝗈𝗆𝗋𝗂𝗇𝗀(𝒱assd,k,H(1+X^(v^),2X^(v^);)).\Phi^{\ast}_{d,k}:\mathcal{QT}^{\mathsf{imm}}_{d,d}(\hat{X},\hat{v};\mathbf{c}\Theta_{\max\geq k};\mathbf{c}\Theta_{\max\geq k})\to\mathsf{Hom_{ring}}\big{(}\mathcal{V}ass_{d,k},\;H^{\ast}(\partial_{1}^{+}\hat{X}(\hat{v}),\partial_{2}^{-}\hat{X}(\hat{v});\mathbb{Z})\big{)}.

In particular, for any generator el𝒱assd,ke_{l}\in\mathcal{V}ass_{d,k}, we get a characteristic element (Φα)(el)Hl(k2)(1+X^(v^),2X^(v^);)(\Phi^{\alpha})^{\ast}(e_{l})\in H^{l(k-2)}(\partial_{1}^{+}\hat{X}(\hat{v}),\partial_{2}^{-}\hat{X}(\hat{v});\mathbb{Z}) which is an invariant of the quasitopy class of α\alpha.

If 2X^(v^)=\partial_{2}^{-}\hat{X}(\hat{v})=\emptyset, then (Φα)(el)(\Phi^{\alpha})^{\ast}(e_{l}) lives in Hl(k2)(X^;)Hl(k2)(1+X^(v^);)H^{l(k-2)}(\hat{X};\mathbb{Z})\approx H^{l(k-2)}(\partial_{1}^{+}\hat{X}(\hat{v});\mathbb{Z}).

If X^\hat{X} is oriented and (n+1)(n+1)-dimensional, using the Poincaré duality in 1+X^(v^)\partial_{1}^{+}\hat{X}(\hat{v}), we produce a homology class 𝒟((Φα)(el))Hnl(k2)(1+X^(v^);)Hnl(k2)(X^;)\mathcal{D}((\Phi^{\alpha})^{\ast}(e_{l}))\in H_{n-l(k-2)}(\partial_{1}^{+}\hat{X}(\hat{v});\mathbb{Z})\approx H_{n-l(k-2)}(\hat{X};\mathbb{Z}) which is again an invariant of the quasitopy class of α\alpha. \diamondsuit

Remark 4.4.

We notice that if (Φα)(el)=0(\Phi^{\alpha})^{\ast}(e_{l})=0, then, by (4.10) and (4.11), all {(Φα)(eq)}ql\{(\Phi^{\alpha})^{\ast}(e_{q})\}_{q\geq l} must be torsion elements in Hq(k2)(1+X^(v^),2X^(v^);)H^{q(k-2)}(\partial_{1}^{+}\hat{X}(\hat{v}),\partial_{2}^{-}\hat{X}(\hat{v});\mathbb{Z}). For an orientable X^\hat{X}, they may be viewed as elements of Hnq(k2)(X^;)H_{n-q(k-2)}(\hat{X};\mathbb{Z}). \diamondsuit

4.10. How to manufacture convex envelops with desired combinatorial tangency patterns

Let us recall one classical construction, leading to the Alexander duality. Let KSdK\subset S^{d} be a CWCW-subcomplex of the dd-sphere. For an element aHp(K;)a\in H_{p}(K;\mathbb{Z}), we denote by caHp+1(Sd,K;)c_{a}\in H_{p+1}(S^{d},K;\mathbb{Z}) the unique element such that (ca)=a\partial_{\ast}(c_{a})=a. The Alexander duality operator 𝒜\mathcal{A}\ell is defined by the formula 𝒜(a)=𝖽𝖾𝖿𝒟(ca)Hdp1(SdK;)\mathcal{A}\ell(a)=_{\mathsf{def}}\mathcal{D}(c_{a})\in H^{d-p-1}(S^{d}\setminus K;\mathbb{Z}), where 𝒟\mathcal{D} is the Poincaré duality operator, the inverse of the operator [Sd][S^{d}]\cap. Pick bHdp1(SdK;)b\in H_{d-p-1}(S^{d}\setminus K;\mathbb{Z}). Then linking number of a,ba,b is defined by the formula 𝗅𝗄(a,b):=𝒜(a),b\mathsf{lk}(a,b):=\langle\mathcal{A}\ell(a),\,b\rangle, where ,\langle\sim,\sim\rangle is the natural pairing between cohomology and homology of dimension np1n-p-1.

The next lemma is instrumental in producing examples of convex envelopes with prescribed combinatorial patterns of their trajectories.

Lemma 4.4.

For any element ω𝛀d]\omega\in\mathbf{\Omega}_{\langle d]}, with the help of the differential (ω)\partial(\omega), given by formula (2.4)(\ref{eq.d+d}), the Alexander duality 𝒜\mathcal{A}\ell produces a cohomology class

θω=𝖽𝖾𝖿𝒜([𝖱¯ω])H|ω|(𝒫d𝐜{ω};),\theta_{\omega}=_{\mathsf{def}}\mathcal{A}\ell([\partial\bar{\mathsf{R}}^{\omega}])\in\;H^{|\omega|^{\prime}}(\mathcal{P}_{d}^{\mathbf{c}\{\omega_{\succ}\}};\mathbb{Z}),

where ω\omega_{\succ} denotes the set of elements of 𝛀d]\mathbf{\Omega}_{\langle d]} that are smaller than ω\omega (so, ω𝐜{ω}\omega\in\mathbf{c}\{\omega_{\succ}\}), and 𝖱¯ω\partial\bar{\mathsf{R}}^{\omega} denotes the algebraic boundary of the cell 𝖱¯ω\bar{\mathsf{R}}^{\omega}.

Proof.

For any ω𝛀d]\omega\in\mathbf{\Omega}_{\langle d]}, take the closed poset ω=𝖽𝖾𝖿{ωω}𝛀d]\omega_{\succ}=_{\mathsf{def}}\{\omega^{\prime}\prec\omega\}\subset\mathbf{\Omega}_{\langle d]} for the role of Θ\Theta in Theorem 2.1. We denote by 𝖱¯dω\bar{\mathsf{R}}_{d}^{\omega} the one-point compactification of the closed cell 𝖱dω𝒫d\mathsf{R}_{d}^{\omega}\subset\mathcal{P}_{d} (the interior (𝖱dω)(\mathsf{R}_{d}^{\omega})^{\circ} of 𝖱dω\mathsf{R}_{d}^{\omega} is an open (d|ω|)(d-|\omega|^{\prime})-ball). Then the differential (ω)\partial(\omega), given by the formula (2.4), represents the (d|ω|1)(d-|\omega|^{\prime}-1)-cycle 𝖱¯dω\partial\bar{\mathsf{R}}_{d}^{\omega} in the chain complex 𝒞(𝒫¯dω;)\mathcal{C}_{\ast}(\bar{\mathcal{P}}_{d}^{\,\omega_{\succ}};\mathbb{Z}) (note that 𝖱¯dω\partial\bar{\mathsf{R}}_{d}^{\omega} is a boundary in 𝒫¯dω\bar{\mathcal{P}}_{d}^{\,\omega_{\succeq}}, but not in 𝒫¯dω\bar{\mathcal{P}}_{d}^{\,\omega_{\succ}} !) and thus defines a nontrivial element [𝖱¯dω]Hd|ω|1(𝒫¯dω;)[\partial\bar{\mathsf{R}}_{d}^{\,\omega}]\in H_{d-|\omega|^{\prime}-1}(\bar{\mathcal{P}}_{d}^{\,\omega_{\succ}};\mathbb{Z}). By the Alexander duality, this element produces a cohomology class θω=𝖽𝖾𝖿𝒜([𝖱¯dω])H|ω|(𝒫d𝐜{ω};).\theta_{\omega}=_{\mathsf{def}}\mathcal{A}\ell([\partial\bar{\mathsf{R}}_{d}^{\omega}])\in H^{|\omega|^{\prime}}(\mathcal{P}_{d}^{\mathbf{c}\{\omega_{\succ}\}};\mathbb{Z}).

Example 4.10.

If d=6d=6, we get the following cohomology classes:

  • θ(121)=𝒜(𝖱¯6(31)𝖱¯6(13)𝖱¯6(2121)+𝖱¯6(1212))H1(𝒫6𝐜{(121)};)\theta_{(121)}=\mathcal{A}\ell\big{(}\bar{\mathsf{R}}_{6}^{(31)}-\bar{\mathsf{R}}_{6}^{(13)}-\bar{\mathsf{R}}_{6}^{(2121)}+\bar{\mathsf{R}}_{6}^{(1212)}\big{)}\in H^{1}(\mathcal{P}_{6}^{\;\mathbf{c}\{(121)_{\succ}\}};\mathbb{Z}),

  • θ(3111)=𝒜(𝖱¯6(411)𝖱¯6(321)+𝖱¯6(312))H2(𝒫6𝐜{(3111)};)\theta_{(3111)}=\mathcal{A}\ell\big{(}\bar{\mathsf{R}}_{6}^{(411)}-\bar{\mathsf{R}}_{6}^{(321)}+\bar{\mathsf{R}}_{6}^{(312)}\big{)}\in H^{2}(\mathcal{P}_{6}^{\mathbf{c}\{(3111)_{\succ}\}};\mathbb{Z}),

  • θ(31)=𝒜(𝖱¯6(4)𝖱¯6(231)+𝖱¯6(321)𝖱¯6(312))H2(𝒫6𝐜{(31)};)\theta_{(31)}=\mathcal{A}\ell\big{(}\bar{\mathsf{R}}_{6}^{(4)}-\bar{\mathsf{R}}_{6}^{(231)}+\bar{\mathsf{R}}_{6}^{(321)}-\bar{\mathsf{R}}_{6}^{(312)}\big{)}\in H^{2}(\mathcal{P}_{6}^{\mathbf{c}\{(31)_{\succ}\}};\mathbb{Z}),

  • θ(1221)=𝒜(𝖱¯6(321)𝖱¯6(141)+𝖱¯6(123))H2(𝒫6𝐜{(1221)};)\theta_{(1221)}=\mathcal{A}\ell\big{(}\bar{\mathsf{R}}_{6}^{(321)}-\bar{\mathsf{R}}_{6}^{(141)}+\bar{\mathsf{R}}_{6}^{(123)}\big{)}\in H^{2}(\mathcal{P}_{6}^{\mathbf{c}\{(1221)_{\succ}\}};\mathbb{Z}). \diamondsuit

Theorem 4.8 below gives a simple recipe for manufacturing traversing flows with the desired number of vv-trajectories of a given combinatorial type ω\omega on compact smooth manifolds XX with boundary. Although the resulting construction is explicit, the topological nature of the pull-back XX is opec due to, paraphrasing Thom [Th], [Th1], “the misterious nature of transversality”.

Theorem 4.8.

For any ω𝛀d]\omega\in\mathbf{\Omega}_{\langle d]}, let θωH|ω|(𝒫d𝐜{ω};)\theta_{\omega}\in H^{|\omega|^{\prime}}(\mathcal{P}_{d}^{\mathbf{c}\{\omega_{\succ}\}};\mathbb{Z}) be the cocycle that takes the value 𝗅𝗄(Z,𝖱¯ω)\mathsf{lk}(Z,\,\partial\bar{\mathsf{R}}^{\omega}) on each |ω||\omega|^{\prime}-dimensional cycle ZZ from H|ω|(𝒫d𝐜{ω};)H_{|\omega|^{\prime}}(\mathcal{P}_{d}^{\mathbf{c}\{\omega_{\succ}\}};\mathbb{Z}). Let YY be an oriented closed and smooth |ω||\omega|^{\prime}-dimensional manifold. We denote by v^\hat{v} a non-vanishing vector field that is tangent to the fibers of the projection ×YY\mathbb{R}\times Y\to Y. Let a map Φ:Y𝒫d𝐜(ω)\Phi:Y\to\mathcal{P}_{d}^{\;\mathbf{c}(\omega_{\succ})} be (d)(\partial\mathcal{E}_{d})-regular in the sense of Definition 3.4 from [K9] (see also (2))777By [K9], Corollary 3.2, such maps form and open and dense set in the space of all smooth maps..

  • Then the natural coupling

    Φ(θω),[Y]=Φ(𝒜(𝖱¯dω)),[Y]\langle\Phi^{\ast}(\theta_{\omega}),\,[Y]\rangle=\langle\Phi^{\ast}(\mathcal{A}\ell(\partial\bar{\mathsf{R}}_{d}^{\omega})),\,[Y]\rangle

    gives the oriented count of the v^\hat{v}-trajectories of the combinatorial type ω\omega in the (|ω|+1)(|\omega|^{\prime}+1)-dimensional convex envelop (×Y,v^)(\mathbb{R}\times Y,\,\hat{v}\big{)} of the pair

    (XΦ=𝖽𝖾𝖿{(u,y)|Φ(y)(u)0},v^)(×Y,v^).\big{(}X_{\Phi}=_{\mathsf{def}}\{(u,y)|\;\Phi(y)(u)\leq 0\},\,\hat{v}\big{)}\subset\big{(}\mathbb{R}\times Y,\,\hat{v}\big{)}.

    The combinatorial types of the v^\hat{v}-trajectories relative to XΦ\partial X_{\Phi} belong to the poset 𝐜{ω}=ω=𝖽𝖾𝖿{ω𝛀d]|ωω}\mathbf{c}\{\omega_{\succ}\}=\omega_{\preceq}=_{\mathsf{def}}\{\omega^{\prime}\in\mathbf{\Omega}_{\langle d]}|\;\omega^{\prime}\succeq\omega\}.

  • If Φ\Phi is transversal to the cell (𝖱dω)(\mathsf{R}_{d}^{\omega})^{\circ} in 𝒫d𝐜{ω}\mathcal{P}_{d}^{\;\mathbf{c}\{\omega_{\succ}\}}, then the number of v^\hat{v}-trajectories in ×Y\mathbb{R}\times Y, whose intersection with XΦ\partial X_{\Phi} has the combinatorial type ω\omega, equals the cardinality of the intersection Φ(Y)(𝖱dω)\Phi(Y)\cap(\mathsf{R}_{d}^{\omega})^{\circ}. Thus the number of such trajectories is greater than or equal to the absolute value |Φ(θω),[Y]||\langle\Phi^{\ast}(\theta_{\omega}),[Y]\rangle|.

Proof.

The argument is similar to the proof of Lemma 4.4. By Theorem 2.1, the homology class of the cycle 𝖱¯dω\partial\bar{\mathsf{R}}_{d}^{\omega} in Hd|ω|1(𝒫¯d{ω};)H_{d-|\omega|^{\prime}-1}(\bar{\mathcal{P}}_{d}^{\{\omega_{\succ}\}};\mathbb{Z}), represented by formula (2.4), is nontrivial since d|ω|1d-|\omega|^{\prime}-1 is the top grading of the differential complex ([ω],)(\mathbb{Z}[\omega_{\succ}],\partial).

The proof amounts to spelling out the nature of Alexander duality 𝒜\mathcal{A}\ell. By its definition, Z,𝒜(𝖱¯dω)=𝗅𝗄(Z,𝖱¯dω)\langle Z,\,\mathcal{A}\ell(\partial\bar{\mathsf{R}}_{d}^{\omega})\rangle=\mathsf{lk}(Z,\partial\bar{\mathsf{R}}_{d}^{\omega}) for any cycle ZZ in 𝒫d𝐜(ω)\mathcal{P}_{d}^{\;\mathbf{c}(\omega_{\succ})} of dimension |ω||\omega|^{\prime}. Thus

Φ(θω),[Y]=θω,Φ([Y])=𝗅𝗄(𝖱¯dω,Φ(Y))=𝖱¯dωΦ(Y).\langle\Phi^{\ast}(\theta_{\omega}),[Y]\rangle=\langle\theta_{\omega},\Phi_{\ast}([Y])\rangle=\mathsf{lk}(\partial\bar{\mathsf{R}}_{d}^{\omega},\,\Phi(Y))=\bar{\mathsf{R}}^{\omega}_{d}\,\circ\,\Phi(Y).

Examining the construction of the space XΦY×X_{\Phi}\subset Y\times\mathbb{R}, given by (𝗂𝖽×Φ)1(d)(\mathsf{id}\times\Phi)^{-1}(\mathcal{E}_{d}), where the hypersurface d\mathcal{E}_{d} was introduced in (2), the latter intersection number gives an oriented count of the v^\hat{v}-trajectories in ×Y\mathbb{R}\times Y of the combinatorial type ω\omega relative to the boundary XΦ\partial X_{\Phi}.

If Φ:Y𝒫d𝐜(ω)\Phi:Y\to\mathcal{P}_{d}^{\;\mathbf{c}(\omega_{\succ})} is transversal to the open cell (𝖱dω)(\mathsf{R}^{\omega}_{d})^{\circ}, then the cardinality of the geometric intersection Φ(Y)(𝖱ω)\Phi(Y)\cap(\mathsf{R}^{\omega})^{\circ} is exactly the total number of v^\hat{v}-trajectories of the combinatorial type ω\omega with respect to XΦ\partial X_{\Phi}. The intersection points from Φ(Y)(𝖱dω)\Phi(Y)\cap(\mathsf{R}_{d}^{\omega})^{\circ} (equivalently, the trajectories of the type ω\omega with respect to XΦ\partial X_{\Phi}) come in two flavors, {,}\{\oplus,\ominus\}, depending on whether the canonicalal normal orientation of the cell (𝖱dω)(\mathsf{R}_{d}^{\omega})^{\circ} in 𝒫d\mathcal{P}_{d} agrees or disagrees with the preferred orientation of the cycle Φ(Y)\Phi(Y). ∎

The next corollary follows directly from Theorem 4.8.

Corollary 4.11.

Take ω=(1,2,,2n,1)\omega=(1,\underbrace{2,\dots,2}_{n},1). Let Φ:Yn𝒫2n+2𝐜{ω}\Phi:Y^{n}\to\mathcal{P}_{2n+2}^{\;\mathbf{c}\{\omega_{\succ}\}} be as in Theorem 4.8.

Then the oriented number of trajectories of the combinatorial type ω\omega in the pull-back XΦn+1×YnX^{n+1}_{\Phi}\subset\mathbb{R}\times Y^{n} equals the linking number of the cycle Φ(Yn)\Phi(Y^{n}) with the (n+1)(n+1)-cycle

𝖱¯2n+2(1221)=𝖱¯2n+2(3221)𝖱¯2n+2(14221)++(1)n1𝖱¯2n+2(12241)+(1)n𝖱¯2n+2(1223).\partial\bar{\mathsf{R}}_{2n+2}^{(12\dots 21)}=\bar{\mathsf{R}}_{2n+2}^{(32\dots 21)}-\bar{\mathsf{R}}_{2n+2}^{(142\dots 21)}+\dots+(-1)^{n-1}\,\bar{\mathsf{R}}_{2n+2}^{(12\dots 241)}+(-1)^{n}\,\bar{\mathsf{R}}_{2n+2}^{(12\dots 23)}.

In particular, the number of ω\omega-tangent trajectories in XΦn+1X^{n+1}_{\Phi} is at least

|𝗅𝗄(Φ(Yn),𝖱¯2n+2(1221))|=|h(θ(1221)),[Y]|.|\,\mathsf{lk}(\Phi(Y^{n}),\,\partial\bar{\mathsf{R}}_{2n+2}^{(12\dots 21)})|=|\langle h^{\ast}(\theta_{(12\dots 21)}),\,[Y]\rangle|.\quad\quad\hfill\diamondsuit

4.11. Comments and questions

The following basic, however, non-trivial question animates many of our previous investigations:

“For a given closed poset Θ𝛀d]\Theta\subset\mathbf{\Omega}_{\langle d]}, what are the restrictions on the topology of compact manifolds XX that admit traversing vv-flows (or their pseudo-envelops/envelops (X^,v^)(\hat{X},\hat{v})) whose tangency patterns avoid Θ\Theta?”

In particular, what smooth topological types XX may arise via classifying maps Φ\Phi (say, as in Theorem 4.8 or Corollary 4.11)? We do not have any problems with manufacturing d\partial\mathcal{E}_{d}-regular maps from a given compact nn-manifold YY to 𝒫d𝐜Θ\mathcal{P}_{d}^{\mathbf{c}\Theta}, d0mod2d\equiv 0\mod 2, and using such maps to produce embeddings α:X×Y\alpha:X\subset\mathbb{R}\times Y and traversing flows on XX that avoid the Θ\Theta-patterns. However, the topological types of such XX’s are beyond our control: we get what we get… The resulting XX is subject to many restrictions, whose nature we do not understand conceptually. Let us sketch just a couple examples which indicate that the restrictions on XX by Θ\Theta can be severe.

In [K10], and [K7], Chapters 2-4, we proposed an answer this question for 33-folds XX. For them, the minimal number gc(X)gc(X) of vv-trajectories of the combinatorial type ω=(1221)\omega=(1221) was introduced as a measure of complexity of a 33-fold XX and was linked directly to the combinatorial complexity c(X)c(X) of their 22-spines. Via this link, gc(X)gc(X) is related to the classification of 33-folds. For instance, a connected compact oriented 33-fold with a simply-connected boundary, which admits a traversing flow that avoids Θ=(1221)\Theta=(1221), is a connected sum of several 33-balls and products S1×S2S^{1}\times S^{2} ([K10], Theorem 3.14). Therefore, no other 33-folds with a simply-connected boundary can admit a convex pseudo-envelop or even a traversing flow that avoids the tangency pattern (1221)(1221).

In a quite different setting, consider a closed hyperbolic (n+1)(n+1)-manifold ZZ. Let XX be obtained from ZZ by deleting a number (n+1)(n+1)-balls. Then no such XX can support a traversing vv-flow that avoids the set of isolated v^\hat{v}-trajectories of combinatorial types {ω𝛀||=n}\{\omega\in\mathbf{\Omega}_{|\sim|^{\prime}=n}\}. Indeed, the positivity of Gromov’s simplicial semi-norm ZΔ\|Z\|_{\Delta} rules out the possibility of such an avoidance (see [AK], Theorem 1). In fact, the positivity of simplicial semi-norms of various homology classes hH(X;)h\in H_{\ast}(X;\mathbb{R}) is the only general mechanism known to us that imposes constraints on the combinatorial tangency types of any generic traversing flow on XX [K11].

Let us conclude with the following remark. Unfortunately, we do not know examples of invariants that distinguish between the quasitopies 𝒬𝒯d,d𝖾𝗆𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{emb}}_{d,d^{\prime}}(\hat{X},\hat{v};\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) and 𝒬𝒯d,d𝗌𝗎𝖻(X^,v^;𝐜Θ;𝐜Θ)\mathcal{QT}^{\mathsf{sub}}_{d,d^{\prime}}(\hat{X},\hat{v};\hfill\break\mathbf{c}\Theta;\mathbf{c}\Theta^{\prime}) of convex envelops and pseudo-envelops. It seems natural to think about invariants that utilize the kk-multiple self-intersection manifolds {Σkα}k2\{\Sigma_{k}^{\alpha^{\partial}}\}_{k\geq 2} of α\alpha^{\partial} (see Remark 4.3). However, for the convex pseudo-envelops, the analogue of the distinguishing map (3.3) in [K9] is trivial: in fact, for a submersion α:XX^\alpha:X\to\hat{X}, the bordism class of the map παp1:ΣkαX^\pi\circ\alpha^{\partial}\circ p_{1}:\Sigma_{k}^{\alpha^{\partial}}\to\hat{X} vanishes, due to arguments as in [K9], Corollary 3.1. As a result, a direct analogue of Proposition 3.6 from [K9] is vacuous in the environment of convex envelops.

Acknowledgment: The author is grateful to the Department of Mathematics of Massachusetts Institute of Technology for many years of hospitality.

References

  • [AK] Alpert, H., Katz, G., Using Simplicial Volume to Count Multi-tangent Trajectories of Traversing Vector Fields, Geometriae Dedicata, 323-338 (2016).
  • [Ar] Arnold, V.I., Spaces of functions with moderate singularities. (Russian) Funktsional. Anal. i Prilozhen. 23 (1989), 1-10; translation in Funct. Anal. Appl. 23 (1990) 169-177.
  • [Ar1] Arnold, V.I., Springer numbers and morsification spaces, J. Algebraic Geom., 1 (2) (1992), 197-214.
  • [Bo] Boardman, J. M., Singularities of Differentiable Maps, Publ. Math. I.H.E.S., 33 (1967), pp. 21-57.
  • [GG] Golubitsky, M., Guillemin, V., Stable Mappings and Their Singularities, Graduate Texts in Mathematics 14, Springer-Verlag, New York Heidelberg Berlin, 1973.
  • [Ha] Hatcher, A., Algebraic Topology, Cambridge University Press, 2002.
  • [Hu] Hu, S.T., Homotopy Theory, Academic Press, New York - London, 1959.
  • [K1] Katz, G., Stratified Convexity and Concavity of Gradient Flows on Manifolds with Boundary, Applied Mathematics, 2014, 5, 2823-2848, (SciRes. http://www.scirp.org/journal/am) (also arXiv:1406.6907v1 [mathGT] (26 June, 2014)).
  • [K2] Katz, G., Traversally Generic & Versal Flows: Semi-algebraic Models of Tangency to the Boundary, Asian J. of Math., vol. 21, No. 1 (2017), 127-168 (arXiv:1407.1345v1 [mathGT] 4 July, 2014)).
  • [K3] Katz, G., The Stratified Spaces of Real Polynomials and Trajectory Spaces of Traversing Flows, JP Journal of Geometry and Topology, Vol. 19, No 2, 2016, 95-160.
  • [K5] Katz G., Flows in Flatland: A Romance of Few Dimensions, Arnold Math. J., DOI 10.1007/s40598-016-0059-1, Springer (2016).
  • [K6] Katz, G., Causal Holography in Application to the Inverse Scattering Problem, Inverse Problems and Imaging J., June 2019, 13(3), 597-633 (arXiv: 1703.08874v1 [Math.GT], 27 Mar 2017).
  • [K7] Katz, G., Morse Theory of Gradient Flows, Concavity and Complexity on Manifolds with Boundary, World Scientific (2019), New Jersey - London - Singapore - Beijing - Shanghai - Hong Kong - Taipei - Chennai - Tokyo, ISBN 978-981-4368-75-9.
  • [K8] Katz, G., Holography of geodesic flows, harmonizing metrics, and billiards’ dynamics, arXiv: 2003.10501 v2 [math.DS] 1 April 2020.
  • [K9] Katz, G., Spaces of polynomials with constrained divisors as Grassmanians for immersions & embeddings, arXiv:2201.02744v1 [math GT] 8 Jan 2022.
  • [K10] Katz, G., Convexity of Morse Stratifications and Spines of 3-Manifolds, JP Journal of Geometry and Topology, vol. .9, No 1 (2009), 1-119.
  • [K11] Katz, G., Complexity of shadows and traversing flows in terms of the simplicial volume, Journal of Topology and Analysis, Vol. 8, No. 3 (2016) 501-543.
  • [KSW1] Katz, G., Shapiro B., Welker, V., Real polynomials with constrained real divisors. I. Fundamental Groups, arXiv:1908.07941 v3 [math.AT] 18 Sep 2020.
  • [KSW2] Katz, G., Shapiro B., Welker, V., Spaces of polynomials and bilinear forms with constrained real divisors, II. (Co)homology & stabilization, arXiv:2112.15205v1 [math. AT] 30 Dec. 2021.
  • [Ke] Kervaire, M., Smooth Homology Spheres and their Fundamental Groups, Transactions of AMS, 144 : 67-72, doi:10.2307/1995269.
  • [KeM] Kervaire, M., Milnor, J., Groups of homotopy spheres: I. Annals of Mathematics, 77 (3): 504-537 (1963).
  • [LS] Lashof, R., Smale, S., Self-intersections of Immersed Manifolds, Journal of Mathematics and Mechanics, Vol. 8, No. 1 (1959), pp. 143-157.
  • [Mi] Milnor, J., Lectures on the h-cobordism theorem, Princeton, New Jersey, Princeton University Press, 1965.
  • [Mor] Morin, B., Formes canoniques des singularities d’une application differentiable, Comptes Rendus Acad. Sci., Paris 260, (1965) 5662-5665, 6503-6506.
  • [Mo] Morse, M. Singular points of vector fields under general boundary conditions, Amer. J. Math. 51 (1929), 165-178.
  • [Pa] Pappas, C., Extensions of codimension one immersions, Trans. of AMS, v. 348, no. 1, August 1996.
  • [P1] Perelman, G., The entropy formula for the Ricci flow and its geometric applications, (2002). arXiv:math.DG/0211159.
  • [P2] Perelman, G., Ricci flow with surgery on three-manifolds, (2003). arXiv:math.DG/0303109.
  • [P3] Perelman, G., Finite extinction time for the solutions to the Ricci flow on certain three-manifolds, (2003). arXiv:math.DG/0307245.
  • [S] Smale, S., The classification of immersions of spheres in Euclidean spaces, Ann. Math. 69, 27-344 (1959). MR0105117 (21:3862).
  • [Th] Thom, R., La classification des immersions (d’après Smale), Séminaire Bourbaki, Exposé 157, 11 pp. (1957) MR0105693 (21:4429).
  • [Th1] Thom, R., Les singularites des applications differentiable, Comm. Math. Helv., 28 (1954), 17-86.
  • [V] Vassiliev, V.A., Complements of Discriminants of Smooth Maps: Topology and Applications, Translations of Mathematical Monographs, vol 98, American Mathematical Society (1994).
  • [W] Wall, C.T.C., Non-Additivity of the Signature, Inventiones Math., 7, 269-274 (1969).
  • [WaX] Wang, G., Xu, Z., The triviality of the 61-stem in the stable homotopy groups of spheres, Annals of Mathematics, 186 (2): 501-580, (2017), arXiv:1601.02184
  • [Wh] Whitehead, G., W., Elements of Homotopy Theory, Springer-Verlag, New York-Heidelberg-Berlin, 1978.