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

The space of sections of a smooth function

Gunnar Carlsson, Benjamin Filippenko
Abstract.

Given a compact manifold XX with boundary and a submersion f:XYf:X\rightarrow Y whose restriction to the boundary of XX has isolated critical points with distinct critical values and where YY is [0,1][0,1] or S1S^{1}, the connected components of the space of sections of ff are computed from π0\pi_{0} and π1\pi_{1} of the fibers of ff. This computation is then leveraged to provide new results on a smoothed version of the evasion path problem for mobile sensor networks: From the time-varying homology of the covered region and the time-varying cup-product on cohomology of the boundary, a necessary and sufficient condition for existence of an evasion path and a lower bound on the number of homotopy classes of evasion paths are computed. No connectivity assumptions are required.

1. Introduction

Given a smooth map f:XYf:X\rightarrow Y of compact manifolds, we are interested in the homotopy type of the space of continuous sections

Γf={δ:YX|fδ=idY}\Gamma f=\{\delta:Y\rightarrow X\,\,|\,\,f\circ\delta=id_{Y}\}

of ff. In the case that ff is a fiber bundle, there is a well-known obstruction theory for constructing sections of ff by moving up the skeleta of a CW-decomposition of YY. If in addition YY is contractible, then Γf\Gamma f is homotopy equivalent to the fiber of ff. The situation when ff is not a fiber bundle has received much less attention in the literature. It has interesting behavior even when YY is contractible.

For Y=[0,1]Y=[0,1] and Y=S1Y=S^{1}, we establish a complete computation (Theorem 1.1) of π0(Γf)\pi_{0}(\Gamma f) for a class of smooth maps f:XYf:X\rightarrow Y that is generic on the boundary X\partial X. We call these tame functions (Definition 2.2). The computation uses π1\pi_{1} and π0\pi_{0} of the fibers of ff. Roughly, a tame function ff on a smooth compact XX with boundary X\partial X is a submersion whose restriction f|Xf|_{\partial X} has isolated critical points with distinct critical values. In Remark 1.3, we propose various extensions of this result, e.g. a computation of πk\pi_{k} of the components of Γf\Gamma f, higher dimensional YY, and other interesting possibilities.

We apply Theorem 1.1 to a smoothed version of the evasion path problem for mobile sensor networks §4. A mobile sensor network is a collection of sensors moving continuously in a bounded domain 𝒟d{\mathcal{D}}\subset\mathbb{R}^{d} such that each sensor can detect objects within a fixed radius. The evasion path problem asks for the identification of intruders that follow a continuous path in the domain 𝒟{\mathcal{D}} that is at all times disjoint from the region CC covered by the sensors. Intruders should be identified from only topological (e.g., no coordinates) information about the mobile sensor network. This problem has been studied in [6] [1] [7] (see §4.1).

We introduce a smoothed version of the evasion path problem in §4.2 which effectively approximates the mobile sensor network version. In Corollary 4.2, we obtain a complete computation of the connected components of the space of evasion paths in terms of the time-varying π0\pi_{0} and π1\pi_{1} of the uncovered region X=𝒟CX={\mathcal{D}}\setminus C. In Theorem 1.4, we establish a necessary and sufficient condition for existence of an evasion path and moreover a lower bound on the number of connected components in terms of time-varying (co)homological information about the covered region CC and its boundary. The cup product plays a crucial role. Note that Theorem 1.4 does not require the time-varying covered region CtC_{t} to be connected, unlike all prior necessary and sufficient conditions for existence of an evasion path. In the connected case, Theorem 1.4 has the simpler form Corollary 1.5.

The precise situation in Theorem 1.1 is as follows. Refer to Figure 1 for examples. Let XX be a smooth compact cobordism between manifolds with boundary X0X_{0} and X1X_{1} (Definition 2.1). Let f:X[0,1]f:X\rightarrow[0,1] be a tame function (Defintion 2.2), i.e. a smooth submersion with f1(i)=Xif^{-1}(i)=X_{i} for i=0,1i=0,1 such that the restriction f|X:X[0,1]f|_{\partial X}:\partial X\rightarrow[0,1] to the boundary X\partial X (not including X0X_{0} and X1X_{1}) has isolated critical points with distinct critical values. (There is a similar story for f:XS1f:X\rightarrow S^{1} where XX is a manifold with boundary). Note that ff is submersive if, for example, Xd×[0,1]X\subset\mathbb{R}^{d}\times[0,1] is a codimension-0 embedding and ff is the projection onto [0,1][0,1], as is the case in the smoothed evasion path problem. A next step in this research is to allow ff to have critical points in the interior of XX, making the conditions CC^{\infty}-generic.

To state the theorem, choose regular values sis_{i} that interleave the critical values tit_{i} of f|Xf|_{\partial X}:

0=s0<t1<s1<t2<<tn<sn=1.0=s_{0}<t_{1}<s_{1}<t_{2}<\cdots<t_{n}<s_{n}=1.

Set Xi=f1(si)X_{i}=f^{-1}(s_{i}) and Xii+1=f1([si,si+1]).X_{i}^{i+1}=f^{-1}([s_{i},s_{i+1}]). There is a diagram of spaces where all maps are inclusions of regular level sets into the regular cobordisms between them

(1) ZX~:=(X0X01X1Xn1nXn).\widetilde{ZX}:=\big{(}X_{0}\hookrightarrow X_{0}^{1}\hookleftarrow X_{1}\hookrightarrow\cdots\hookrightarrow X_{n-1}^{n}\hookleftarrow X_{n}\big{)}.

Since each Xii+1X_{i}^{i+1} contains exactly 11 critical point of f|Xf|_{\partial X} on the boundary, we can construct a gradient-like vector field whose flow deformation retracts Xii+1X_{i}^{i+1} onto either XiX_{i} or Xi+1X_{i+1} depending on whether the outward pointing normal vector η\eta at the boundary critical point in Xii+1X_{i}^{i+1} satisfies df(η)>0df(\eta)>0 or df(η)<0df(\eta)<0 (by submersivity of ff, we have df(η)0df(\eta)\neq 0). We keep track of this information via the assignment

(2) +(i):={iif df(η)>0i+1if df(η)<0,\partial^{+}(i):=\begin{cases}i&\text{if }df(\eta)>0\\ i+1&\text{if }df(\eta)<0,\end{cases}

for i=0,,n1i=0,\ldots,n-1. The inclusion X+(i)Xii+1X_{\partial^{+}(i)}\hookrightarrow X_{i}^{i+1} is a homotopy equivalance. After applying π0\pi_{0} to the diagram (1), one of the induced maps going into π0(Xii+1)\pi_{0}(X_{i}^{i+1}) is a bijection, so we obtain a diagram where the arrows point either left or right,

π0(ZX):=(π0(X0)π0(X1)π0(Xn)).\pi_{0}(ZX):=\big{(}\pi_{0}(X_{0})\leftrightarrow\pi_{0}(X_{1})\leftrightarrow\cdots\leftrightarrow\pi_{0}(X_{n})\big{)}.

The following theorem is proved in §3.2; see Theorems 3.14, 3.15 for the precise statements.

Theorem 1.1.

  1. (i)

    There is a surjection

    Π0:π0(Γf)limπ0(ZX)\Pi_{0}:\pi_{0}(\Gamma f)\rightarrow\varprojlim\pi_{0}(ZX)

    with fiber over Ψlimπ0(ZX)\Psi\in\varprojlim\pi_{0}(ZX) characterized as follows. Let 𝔟Γf\mathfrak{b}\in\Gamma f such that Π0(𝔟)=Ψ\Pi_{0}(\mathfrak{b})=\Psi. Then Π01(Ψ)\Pi_{0}^{-1}(\Psi) is naturally in bijection with the orbits of an action of the group i=0nπ1(Xi,𝔟(si))\prod_{i=0}^{n}\pi_{1}(X_{i},\mathfrak{b}(s_{i})) on the set i=0n1π1(X+(i),𝔟(s+(i)))\prod_{i=0}^{n-1}\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)})).

  2. (ii)

    Let XX be a smooth compact manifold with boundary equipped with a submersion f:XS1f:X\rightarrow S^{1} whose restriction to the boundary f|X:XS1f|_{\partial X}:\partial X\rightarrow S^{1} has isolated critical points with distinct critical values. Define the diagram ZX~\widetilde{ZX} as in (1) with the additional identity map X0=XnX_{0}=X_{n}, and similarly define π0(ZX)\pi_{0}(ZX). Then the statements in (i) hold.

Refer to caption Refer to caption(a) (b)
Refer to caption (c)
Figure 1. The 33-dimensional white cobordisms XX are embedded in the grey D2×[0,1]D^{2}\times[0,1] and the projections ff to [0,1][0,1] are tame. The wavy shaded region in (c)(c) is not in XX. The critical points of the restriction f|X:X[0,1]f|_{\partial X}:\partial X\rightarrow[0,1] to the boundary X\partial X are the red dots with critical values tit_{i}. Sections of ff are shown in blue, green, and, orange.
Remark 1.2.

Applying Theorem 1.1 to the examples in Figure 1 produces the following.

In (a)(a), XX is an embedded solid pair of pants and the sections Γf\Gamma f of the projection f:X[0,1]f:X\rightarrow[0,1] have two connected components |π0(Γf)|=2|\pi_{0}(\Gamma f)|=2, represented by the green and blue sections.

In (b)(b), XX is an embedded solid cylinder and there are no sections, Γf=\Gamma f=\emptyset.

In (c), XX is a cobordism between a single annulus and two annuli. There are infinitely many connected components of the sections |π0(Γf)|=|\pi_{0}(\Gamma f)|=\infty. Indeed, the three sections pictured are not fiberwise homotopic, and one can modify the orange section to wrap around the bottom arm of the wavy region any integral number of times, producing a countable collection of non-fiberwise homotopic sections.

Remark 1.3.

We propose the following extensions of these results to a general theory of sections of CC^{\infty}-generic smooth maps f:XYf:X\rightarrow Y.

  1. (i)

    There is a generalization of Theorem 1.1 to a computation of πk(Γf,b)\pi_{k}(\Gamma f,b) for any basepoint bΓfb\in\Gamma f and k1k\geq 1. These higher πk\pi_{k} can be addressed with a fiberwise version of the unstable Adams spectral sequence of Bousfield-Kan. This is currently being pursued by Wyatt Mackey.

  2. (ii)

    What happens when we allow ff to have critical points in the interior of XX? Answering this will complete the dimY=1\dim Y=1 story for CC^{\infty}-generic ff (isolated critical points with distinct critical values).

  3. (iii)

    Is there a sheaf theoretic interpretation of these results? The regular level sets in the theorem are homotopy equivalent to preimages of small open neighborhoods, so the diagram ZX~\widetilde{ZX} in (1) comes from an open covering of [0,1][0,1] and inclusions of intersections of those open sets. This suggests replacing π0(ZX)\pi_{0}(ZX) with the co-presheaf Uπ0(f1(U))U\mapsto\pi_{0}(f^{-1}(U)) on [0,1][0,1].

  4. (iv)

    Generalize Theorem 1.1 to higher dimensions dimY>1\dim Y>1. We expect that the π1\pi_{1}-actions will generalize to πdimY\pi_{\dim Y}-actions.

The Evasion Path Problem: We apply Theorem 1.1 to obtain new results on the evasion path problem in applied topology: Theorem 1.4 and Corollary 1.5. The evasion path problem is summarized as follows; see §4 for a detailed description. Given a collection of continuous sensors 𝒮={γ:[0,1]𝒟}{\mathcal{S}}=\{\gamma:[0,1]\rightarrow{\mathcal{D}}\} moving in a bounded domain 𝒟d{\mathcal{D}}\subset\mathbb{R}^{d} that detect objects within some fixed radius of γ(t)\gamma(t) in d\mathbb{R}^{d}, an evasion path is a continuous intruder δ:[0,1]𝒟\delta:[0,1]\rightarrow{\mathcal{D}} that avoids detection by the sensors for the whole time interval I=[0,1]I=[0,1]. Let Ct𝒟C_{t}\subset{\mathcal{D}} denote the region covered by the sensors at time tIt\in I, and set

Xt=𝒟Ct.X_{t}={\mathcal{D}}\setminus C_{t}.

Then evasion paths δ\delta are sections of the function

f:X:=tIXt×{t}d×II.f:X:=\bigcup_{t\in I}X_{t}\times\{t\}\subset\mathbb{R}^{d}\times I\longrightarrow I.

Let Γf\Gamma f denote the space of evasion paths.

The sensor ball evasion path problem asks for a criterion that determines whether or not an evasion path exists and that is based only on homological information about the covered region CtC_{t}. In practice, one imagines that we can understand the topology of CtC_{t} since it is the region covered by the sensors. For example, if sensors can detect overlaps of their sensed regions, then Čech cohomology of CtC_{t} can be computed. Versions of this problem have been studied in [6] [1] [7]; see 4.1.

We consider an idealized version of the evasion path problem; see §4.2. Roughly, we smooth C=tICt×{t}d×IC=\bigcup_{t\in I}C_{t}\times\{t\}\subset\mathbb{R}^{d}\times I into a smooth cobordism of manifolds with boundary embedded in d×I\mathbb{R}^{d}\times I that closely approximates the region covered by the sensors and whose projection CIC\rightarrow I is tame.

Let BB denote the boundary of CC, except for the interior of the fibers over 0 and 11. Note that both BB and CC have associated diagrams ZB~\widetilde{ZB} and ZC~\widetilde{ZC} defined in the same way as ZX~\widetilde{ZX} in (1).

The precise statement of Theorem 1.4 is given in Theorem 4.5. In Example 4.9, Theorem 1.4 is applied to example (a) from Figure 1. See Remark 4.10 for the d=0,1d=0,1 cases.

Theorem 1.4.

Assume d2d\geq 2. There is a surjection π0(Γf)limHomkalgebra(H0(ZX~;k),k)\pi_{0}(\Gamma f)\rightarrow\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k). In particular, an evasion path exists (i.e. Γf\Gamma f is nonempty) if and only if limHomkalgebra(H0(ZX~;k),k)\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k) is nonempty, and the cardinality of π0(Γf)\pi_{0}(\Gamma f) is bounded from below by the cardinality of the inverse limit.

Assume that the projection CIC\rightarrow I does not have any local minima or local maxima except over 0,1I0,1\in I, which holds if sensors are not created or destroyed in time. Then the zigzag diagram of kk-algebras H0(ZX~;k)H^{0}(\widetilde{ZX};k) is determined up to isomorphism by the zigzag diagram of kk-algebras H0(ZB~;k)H^{0}(\widetilde{ZB};k), the map +\partial^{+} defined in (2), and an Alexander duality isomorphism of H0H^{0} of the regular fibers of BB with Hd1H_{d-1} of their complements BcB^{c} as well as maps on Hd1H_{d-1} induced fiberwise by inclusion CBcC\rightarrow B^{c}.

Corollary 1.5.

Assume that CtC_{t} is connected for all tIt\in I. Then there is a surjection π0(Γf)limHomkalgebra(H0(ZB~;k),k)\pi_{0}(\Gamma f)\rightarrow\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZB};k),k). In particular, an evasion path exists (i.e. Γf\Gamma f is nonempty) if and only if limHomkalgebra(H0(ZB~;k),k)\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZB};k),k) is nonempty, and the cardinality of π0(Γf)\pi_{0}(\Gamma f) is bounded from below by the cardinality of the inverse limit.

Proof.

This follows immediately from Corollary 4.2, Proposition 4.4, and Proposition 4.7. ∎

Remark 1.6.

The following heuristic suggests that in applications to sensor networks it is enough to compute limHomkalgebra(H0(ZX~;k),k)\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k) to understand all evasion paths that an intruder is likely to take. The fibers of the surjection π0(Γf)limHomkalgebra(H0(ZX~;k),k)\pi_{0}(\Gamma f)\rightarrow\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k) are described by π1\pi_{1} of the regular fibers of the map XIX\rightarrow I, as in Theorem 1.1. An intruder is randomly sampling from the space of paths, and so is unlikely to wrap around nontrivial loops in the regular fibers.

Funding. This material is based upon work supported by the National Science Foundation under Award No. 1903023.

2. Tame functions on cobordisms of manifolds with boundary

Definition 2.1.

A compact cobordism of manifolds with boundary is a compact (d+1)(d+1)-dimensional manifold XX with boundary and corners111For 0kn0\leq k\leq n, the kk-stratum kX\partial_{k}X of a nn-dimensional manifold XX with boundary and corners consists of those points xXx\in X around which there is a smooth chart to [0,)k×nk[0,\infty)^{k}\times\mathbb{R}^{n-k} that identifies yy with a point in {0}k×nk\{0\}^{k}\times\mathbb{R}^{n-k}. For a cobordism XX of manifolds with boundary, the highest nonempty stratum is 2X=B=X0X1\partial_{2}X=\partial B=\partial X_{0}\sqcup\partial X_{1}. The 11-stratum 1X\partial_{1}X is the union of the interiors of X1,B,X_{1},B, and X2X_{2}, which together with the 22-stratum forms the full boundary X\partial X. The 0-stratum 0X\partial_{0}X is the interior of XX. whose boundary decomposes as a union

X=XBX+\partial X=X_{-}\cup B\cup X_{+}

where BB and each X±X_{\pm} is an embedded dd-dimensional manifold with boundary and such that the following properties hold:

  • The intersection XX+=X_{-}\cap X_{+}=\emptyset is empty. We say that XX is a cobordism between XX_{-} and X+X_{+},

  • X±=X±B,\partial X_{\pm}=X_{\pm}\cap B,

  • BB is a cobordism between the closed manifolds X\partial X_{-} and X+,\partial X_{+}, i.e., B=XX+.\partial B=\partial X_{-}\sqcup\partial X_{+}.

The following notion of a tame function, as well as the constructions we perform with them in this section, is inspired by the Morse theory on manifolds with boundary; see [2][3][4][8][9][10][11][12] and Remark 2.3.

Definition 2.2.

A tame function f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] on a compact cobordism of manifolds with boundary is a smooth function satisfying the following conditions:

  • The critical points of the restriction f|B:B[s,s+]f|_{B}:B\rightarrow[s_{-},s_{+}] are isolated and have distinct critical values in (s,s+)(s_{-},s_{+}),

  • ff is submersive,

  • X=f1(s)X_{-}=f^{-1}(s_{-}) and X+=f1(s+)X_{+}=f^{-1}(s_{+}).

Remark 2.3.

Definition 2.2 does not require any nondegeneracy condition on the critical points of f|B:B[s,s+]f|_{B}:B\rightarrow[s_{-},s_{+}], so f|Bf|_{B} does not have to be a Morse function in the sense of [12, Def. 2.3]. In this way, our definition is more general than the Morse condition. On the other hand, we do not allow ff itself to have critical points. Note that for the projection f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] from a codimension-0 submanifold Xd×[s,s+]X\subset\mathbb{R}^{d}\times[s_{-},s_{+}], as is the situation in the smoothed evasion path problem (see §4), the map ff is submersive.

There are two types of critical points of a tame function on the boundary.

Definition 2.4.

Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function, pBp\in B a critical point of f|B:B[s,s+]f|_{B}:B\rightarrow[s_{-},s_{+}], and ηTpX\eta\in T_{p}X an outward pointing vector. Then pp is type N if df(η)<0df(\eta)<0 and type D if df(η)>0.df(\eta)>0.

Note that df(η)=0df(\eta)=0 is impossible since it would imply that pp is a critical point of ff.

2.1. Local constructions

The following constructions are local in the sense that for a tame function f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] and t[s,s+t\in[s_{-},s_{+}], the constructions apply in the preimage f1([tϵ,t+ϵ])f^{-1}([t-\epsilon,t+\epsilon]) for ϵ>0\epsilon>0 small enough.

To construct sections of a tame function that pass a critical value, we make use of the flow lines of the gradient-like vector field ξ\xi on XX constructed in the following.

Proposition 2.5.

Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function.

If f|Bf|_{B} has only type D critical points, then there exists a smooth vector field ξ\xi on XX with the following properties:

  1. (i)

    df(ξ)=1df(\xi)=-1 on all of XX.

  2. (ii)

    For all222For pXp\in X_{-}, the vector field ξ\xi constructed in the proof of Proposition 2.5 is outward pointing on Int(X)Int(X_{-}), and on X\partial X_{-} it is outward pointing with respect to XX_{-} and inward pointing with respect to BB. pXXp\in\partial X\setminus X_{-}, the vector ξpTpX\xi_{p}\in T_{p}X is inward pointing.

If f|Bf|_{B} has only type N critical points, then there exists a smooth vector field ξ\xi on XX with the following properties:

  1. (i)

    df(ξ)=1df(\xi)=1 on all of XX.

  2. (ii)

    For all pXX+p\in\partial X\setminus X_{+}, the vector ξpTpX\xi_{p}\in T_{p}X is inward pointing.

Moreover, given any smooth section 𝔟:[s,s+]X\mathfrak{b}:[s_{-},s_{+}]\rightarrow X of ff (i.e.  f𝔟(t)=tf\circ\mathfrak{b}(t)=t) such that 𝔟(t)B\mathfrak{b}(t)\not\in B for all t[s,s+]t\in[s_{-},s_{+}], the vector field ξ\xi can be chosen such that

ddt𝔟(t)={ξ|𝔟(t)if type Dξ|𝔟(t)if type N.\frac{d}{dt}\mathfrak{b}(t)=\begin{cases}-\xi|_{\mathfrak{b}(t)}&\text{if }\text{type D}\\ \xi|_{\mathfrak{b}(t)}&\text{if }\text{type N}.\end{cases}
Proof.

We consider the case of type D critical points; the type N case is symmetric.

It suffices to construct ξ\xi locally in an open neighborhood of every point pXp\in X and then sum up these local vector fields with a partition of unity. For the local construction, let pXp\in X and for now assume that if pBp\in B then pp is a regular point of f|Bf|_{B}. Then by the implicit function theorem there exists an open neighborhood U(p)XU(p)\subset X of pp and a smooth chart φ:U(p)V\varphi:U(p)\xrightarrow{\sim}V such that φ(p)=(f(p),0,,0)V\varphi(p)=(f(p),0,\ldots,0)\in V where VV is an open subset of one of the following spaces depending on where pp sits on XX, and in such a way that

fφ1:V[s,s+]f\circ\varphi^{-1}:V\rightarrow[s_{-},s_{+}]

is the projection onto the first coordinate:

  • If pInt(X)p\in Int(X), then VdimXV\subset\mathbb{R}^{\dim X},

  • If pInt(X)p\in Int(X_{-}), then V[s,)×dimX1V\subset[s_{-},\infty)\times\mathbb{R}^{\dim X-1},

  • If pInt(X+)p\in Int(X_{+}), then V(,s+]×dimX1V\subset(-\infty,s_{+}]\times\mathbb{R}^{\dim X-1},

  • If pInt(B)p\in Int(B), then V×([0,)×dimX2)V\subset\mathbb{R}\times([0,\infty)\times\mathbb{R}^{\dim X-2}),

  • If pXp\in\partial X_{-}, then V[s,)×([0,)×dimX2)V\subset[s_{-},\infty)\times([0,\infty)\times\mathbb{R}^{\dim X-2}),

  • If pX+p\in\partial X_{+}, then V(,s+]×([0,)×dimX2)V\subset(-\infty,s_{+}]\times([0,\infty)\times\mathbb{R}^{\dim X-2}).

In all cases above, the constant vector field

(1,1,0,,0)(-1,1,0,\ldots,0)

on VV pulls back through the diffeomorphism φ\varphi to a vector field ξ\xi on U(p)U(p) satisfying df(ξ)=1df(\xi)=-1, and moreover it is inward pointing along all points pXX=Int(B)X+p\in\partial X\setminus X_{-}=Int(B)\cup X_{+}, as required.

It remains to consider a critical point pBp\in B of type DD. By definition of tame function, pp is in Int(B)=B(XX+)Int(B)=B\setminus(\partial X_{-}\cup\partial X_{+}). Consider a neighborhood U~(p)X\tilde{U}(p)\subset X of pp and a coordinate chart U~(p)V[0,)×dimX1\tilde{U}(p)\xrightarrow{\sim}V\subset[0,\infty)\times\mathbb{R}^{\dim X-1} that sends pp to 0. Then the constant vector field (1,0,,0)(1,0,\ldots,0) pulls back to a vector field ξ\xi on U~(p)\tilde{U}(p) that is inward pointing, and hence dpf(ξp)<0d_{p}f(\xi_{p})<0 since pp is type D. Then df(ξ)<0df(\xi)<0 in a smaller open neighborhood U(p)U~(p)U(p)\subset\tilde{U}(p). Hence the vector field ξ/|df(ξ)|\xi/|df(\xi)| satisfies (i) and (ii) on U(p)U(p), as required.

Consider now the final statement of the proposition where we are given a smooth section 𝔟\mathfrak{b} that is disjoint from BB. For p𝔟(t)p\neq\mathfrak{b}(t) for all tt, choose the neighborhood U(p)U(p) to be disjoint from the image of 𝔟\mathfrak{b} and perform the construction as above. Suppose p=𝔟(t)p=\mathfrak{b}(t) for some tt. Choose U(p)U(p) to be disjoint from BB. Define ξ|𝔟(t)=𝔟(t)\xi|_{\mathfrak{b}(t)}=-\mathfrak{b}^{\prime}(t) for all tt such that 𝔟(t)U(p)\mathfrak{b}(t)\in U(p). Then df(ξ|𝔟(t))=(f𝔟)(t)=1df(\xi|_{\mathfrak{b}(t)})=-(f\circ\mathfrak{b})^{\prime}(t)=-1. Extend ξ\xi over U(p)U(p) so that df(ξ)=1df(\xi)=-1 on U(p)U(p). ∎

Using the flow of ξ\xi, we now construct a deformation retraction of XX onto either XX_{-} or X+X_{+} that move the values of f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] at constant speed ±1\pm 1 along [s,s+][s_{-},s_{+}]. In particular, fibers get mapped to fibers at all times throughout the deformation retraction. This deformation retraction is a workhorse that is used throughout.

Lemma 2.6.

Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function.

If all critical points of f|Bf|_{B} are type DD, then XX deformation retracts onto X=f1(s)X_{-}=f^{-1}(s_{-}). Moreover, there is a deformation retraction

H:X×[0,s+s]XH:X\times[0,s_{+}-s_{-}]\rightarrow X

that moves the fibers of ff along [s,s+][s_{-},s_{+}] at constant speed 1-1 until they reach XX_{-} and stop; precisely, HH satisfies

f(H(x,t))=max{s,f(x)t}f(H(x,t))=\max\{s_{-},\,\,f(x)-t\}

and

H(H(x,t),Δ)=H(x,t+Δ)H(H(x,t),\Delta)=H(x,t+\Delta)

for all Δ\Delta. Moreover, given any smooth section 𝔟:[s,s+]X\mathfrak{b}:[s_{-},s_{+}]\rightarrow X of ff (i.e.  f𝔟(t)=tf\circ\mathfrak{b}(t)=t) disjoint from BB, the deformation retraction HH can be chosen so that

H(𝔟(t),Δ)=𝔟(tΔ)H(\mathfrak{b}(t),\Delta)=\mathfrak{b}(t-\Delta)

for all t[s,s+]t\in[s_{-},s_{+}] and 0Δts0\leq\Delta\leq t-s_{-}.

Symmetrically, if all critical points of f|Bf|_{B} are type NN, then there is a deformation retraction H:X×[0,s+s]XH:X\times[0,s_{+}-s_{-}]\rightarrow X of XX onto X+X_{+} satisfying

f(H(x,t))=min{s+,f(x)+t}f(H(x,t))=\min\{s_{+},\,\,f(x)+t\}

and

H(H(x,t),Δ)=H(x,t+Δ)H(H(x,t),\Delta)=H(x,t+\Delta)

for all Δ\Delta. Moreover, given any smooth section 𝔟:[s,s+]X\mathfrak{b}:[s_{-},s_{+}]\rightarrow X of ff disjoint from BB, the deformation retraction HH can be chosen so that

H(𝔟(t),Δ)=𝔟(t+Δ)H(\mathfrak{b}(t),\Delta)=\mathfrak{b}(t+\Delta)

for all t[s,s+]t\in[s_{-},s_{+}] and 0Δs+t0\leq\Delta\leq s_{+}-t.

Proof.

Assume that all critical points of f|Bf|_{B} are type D; the type N case is symmetric. Let ξ\xi be a vector field on XX having the properties guaranteed by Proposition 2.5.

We claim that, for xXx\in X, the flow φt(x)\varphi_{t}(x) of ξ\xi exists for all 0tf(x)s0\leq t\leq f(x)-s_{-}. Indeed, let γx:[0,r]X\gamma_{x}:[0,r]\rightarrow X be the flow line of ξ\xi starting at

γx(0)=x.\gamma_{x}(0)=x.

We have ddt(fγx)=df(ξ)=1\frac{d}{dt}(f\circ\gamma_{x})=df(\xi)=-1 everywhere in the domain of γx\gamma_{x} and hence

fγx(t)=f(x)t.f\circ\gamma_{x}(t)=f(x)-t.

Since ξ\xi is inward pointing along the boundary of XX except on XX_{-}, it follows that the flow line γx(t)\gamma_{x}(t) can be extended to larger tt as long as γx(t)X\gamma_{x}(t)\not\in X_{-}, and if γx(t)X\gamma_{x}(t)\in X_{-} then the flow lines stops. Since γx(t)X\gamma_{x}(t)\in X_{-} if and only if s=fγx(t)=f(x)ts_{-}=f\circ\gamma_{x}(t)=f(x)-t, it follows that r=f(x)sr=f(x)-s_{-}. That is, γx\gamma_{x} is defined on the interval [0,f(x)s][0,f(x)-s_{-}]. And, indeed, the flow is given by

φt(x)=γx(t) for all 0tf(x)s.\varphi_{t}(x)=\gamma_{x}(t)\text{ for all }0\leq t\leq f(x)-s_{-}.

The deformation retraction HH is defined using the flow φ\varphi of ξ\xi,

H:X×[0,s+s]\displaystyle H:X\times[0,s_{+}-s_{-}] X\displaystyle\rightarrow X
(x,t)\displaystyle(x,t) {φt(x)if 0tf(x)sφf(x)s(x)if f(x)sts+s.\displaystyle\mapsto\begin{cases}\varphi_{t}(x)&\text{if }0\leq t\leq f(x)-s_{-}\\ \varphi_{f(x)-s_{-}}(x)&\text{if }f(x)-s_{-}\leq t\leq s_{+}-s_{-}.\\ \end{cases}

Indeed, HH is a deformation retraction of XX onto XX_{-} that satisfies the claimed properties.

Given a smooth section 𝔟:[s,s+]X\mathfrak{b}:[s_{-},s_{+}]\rightarrow X of ff, we choose the vector ξ\xi to satisfy 𝔟(t)=ξ\mathfrak{b}^{\prime}(t)=-\xi, as is made possible by Proposition 2.5. The flow of ξ\xi is everywhere tangent to 𝔟\mathfrak{b}. Since also f(φΔ(𝔟(t)))=tΔf(\varphi_{\Delta}(\mathfrak{b}(t)))=t-\Delta, it follows that φΔ(𝔟(t))=𝔟(tΔ)\varphi_{\Delta}(\mathfrak{b}(t))=\mathfrak{b}(t-\Delta). Hence, for t[s,s+]t\in[s_{-},s_{+}] and 0Δts0\leq\Delta\leq t-s_{-}, we have H(𝔟(t),Δ)=φΔ(𝔟(t))=𝔟(tΔ)H(\mathfrak{b}(t),\Delta)=\varphi_{\Delta}(\mathfrak{b}(t))=\mathfrak{b}(t-\Delta), as claimed. ∎

We also make use of the classical statement that, near a regular value tt of f|B:B[s,s+]f|_{B}:B\rightarrow[s_{-},s_{+}], XX is diffeomorphic to the fiber f1(t)f^{-1}(t) times an interval. The precise statement in our setting is the following.

Lemma 2.7.

Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function (in particular a submersion) such that its restriction f|B:B[s,s+]f|_{B}:B\rightarrow[s_{-},s_{+}] also is a submersion. Then, for every t[s,s+]t\in[s_{-},s_{+}], XX is ‘fiberwise diffeormorphic’ to the trivial cobordism [s,s+]×f1(t)[s_{-},s_{+}]\times f^{-1}(t). Precisely, this means that there exists a diffeomorphism

φ:X[s,s+]×f1(t)\varphi:X\xrightarrow{\cong}[s_{-},s_{+}]\times f^{-1}(t)

such that

f=pr[s,s+]φ,f=pr_{[s_{-},s_{+}]}\circ\varphi,

where pr[s,s+]:[s,s+]×f1(t)[s,s+]pr_{[s_{-},s_{+}]}:[s_{-},s_{+}]\times f^{-1}(t)\rightarrow[s_{-},s_{+}] is projection. Moreover, the restriction φ|f1(t):f1(t){t}×f1(t)\varphi|_{f^{-1}(t)}:f^{-1}(t)\rightarrow\{t\}\times f^{-1}(t) is the identity on f1(t)f^{-1}(t).

Proof.

The proof is standard. In summary, construct a gradient-like vector field ξ\xi such that df(ξ)=1df(\xi)=-1 and ξ\xi is tangent to the boundary BB, outward pointing along XX_{-}, and inward pointing along X+X_{+} (using the same method as the proof of Proposition 2.5, or see the proof of [9, Lem. 3.1]). Then build the diffeomorphism from the flow of ξ\xi similarly to Lemma 2.6 (see also the classical reference [12, Thm. 3.4]). ∎

2.2. Zigzag discretization

Let f:XI=[0,1]f:X\rightarrow I=[0,1] be a tame function. Denote the critical values of f|Bf|_{B} by

t1<t2<<tnt_{1}<t_{2}<\cdots<t_{n}

and choose interleaving values

0=s0<t1<s1<t2<<sn1<tn<sn=1.0=s_{0}<t_{1}<s_{1}<t_{2}<\cdots<s_{n-1}<t_{n}<s_{n}=1.

Consider the subspaces of XX given by the preimages

Xi\displaystyle X_{i} :=f1(si), for 0in,\displaystyle:=f^{-1}(s_{i}),\text{ for }0\leq i\leq n,
Xii+1\displaystyle X_{i}^{i+1} :=f1([si,si+1]), for 0in1.\displaystyle:=f^{-1}([s_{i},s_{i+1}]),\text{ for }0\leq i\leq n-1.

Since the sis_{i} are regular values of ff and f|Bf|_{B}, the XiX_{i} are manifolds with boundary Xi=XiB\partial X_{i}=X_{i}\cap B and Xii+1X_{i}^{i+1} is a cobordism between XiX_{i} and Xi+1X_{i+1}.

These cobordisms and the inclusions of their boundaries fit together into a zigzag diagram of manifolds

ZX~:=(X0ηX01X01ηX01+X1ηX12X12ηX12+ηXn1nXn1nηXn1n+Xn).\widetilde{ZX}:=\bigg{(}X_{0}\xhookrightarrow{\eta_{X_{0}^{1}}^{-}}X_{0}^{1}\xhookleftarrow{\eta_{X_{0}^{1}}^{+}}X_{1}\xhookrightarrow{\eta_{X_{1}^{2}}^{-}}X_{1}^{2}\xhookleftarrow{\eta_{X_{1}^{2}}^{+}}\cdots\xhookrightarrow{\eta_{X_{n-1}^{n}}^{-}}X_{n-1}^{n}\xhookleftarrow{\eta_{X_{n-1}^{n}}^{+}}X_{n}\bigg{)}.

Each Xii+1X_{i}^{i+1} for i=0,,n1i=0,\ldots,n-1 contains exactly 11 critical point of f|Bf|_{B}. Suppose this critical point is type DD. It follows from Lemma 2.6 that the inclusion ηXii+1\eta_{X_{i}^{i+1}}^{-} is an inclusion of a deformation retract. Let H:Xii+1×[0,si+1si]Xii+1H:X_{i}^{i+1}\times[0,s_{i+1}-s_{i}]\rightarrow X_{i}^{i+1} be a deformation retraction onto XiX_{i} with the properties provided by the lemma. Then we have

H(,si+1si)ηXii+1=idXi.H(-,s_{i+1}-s_{i})\circ\eta_{X_{i}^{i+1}}^{-}=id_{X_{i}}.

There is a composite map

βi,i+1:=H(,si+1si)ηXii+1+:Xi+1Xi.\beta_{i,i+1}:=H(-,s_{i+1}-s_{i})\circ\eta_{X_{i}^{i+1}}^{+}:X_{i+1}\rightarrow X_{i}.

If the critical point in Xii+1X_{i}^{i+1} is instead type N, then in the same way a map βi,i+1:XiXi+1\beta_{i,i+1}:X_{i}\rightarrow X_{i+1} is obtained. When we do not wish to specify the type of the critical point, we write this map with arrows pointing both ways

βi,i+1:XiXi+1.\beta_{i,i+1}:X_{i}\leftrightarrow X_{i+1}.
Remark 2.8.

The double arrow \leftrightarrow indicates a map either to the right or to the left, not both ways.

To disambiguate when desired, we define the following mappings which record which side of Xii+1X_{i}^{i+1} it deformation retracts onto, and hence determines the direction of βi,i+1\beta_{i,i+1}. Precisely, for i=0,,n1i=0,\ldots,n-1, define

+(i):={iif ti is type-Di+1if ti is type-N,\partial^{+}(i):=\begin{cases}i&\text{if }t_{i}\text{ is type-}D\\ i+1&\text{if }t_{i}\text{ is type-}N,\end{cases}

and also for notational convenience define the opposite mapping

(i):={i+1if ti is type-Diif ti is type-N.\partial^{-}(i):=\begin{cases}i+1&\text{if }t_{i}\text{ is type-}D\\ i&\text{if }t_{i}\text{ is type-}N.\end{cases}

Then, for both type D and type N critical points, the double arrow denoting βi,i+1\beta_{i,i+1} points in the direction

(3) βi,i+1:X(i)X+(i).\beta_{i,i+1}:X_{\partial^{-}(i)}\rightarrow X_{\partial^{+}(i)}.

The diagram ZX~\widetilde{ZX} is then the top ‘zigzag’ of the larger diagram

X01{X_{0}^{1}}{\,\,\,\cdots\,\,\,}Xn1n{X_{n-1}^{n}}X0{X_{0}}X1{X_{1}}{\cdots}Xn1{X_{n-1}}Xn.{X_{n}.}β0,1\scriptstyle{\beta_{0,1}}β1,2\scriptstyle{\beta_{1,2}}βn2,n1\scriptstyle{\beta_{n-2,n-1}}βn1,n\scriptstyle{\beta_{n-1,n}}

It is often more convenient to work with the bottom row, which we denote by

(4) ZX:=(X0β0,1X1β1,2βn1,nXn).ZX:=(X_{0}\xleftrightarrow{\beta_{0,1}}X_{1}\xleftrightarrow{\beta_{1,2}}\cdots\xleftrightarrow{\beta_{n-1,n}}X_{n}).

The diagram ZXZX is equivalent to ZX~\widetilde{ZX} ‘up to homotopy,’ in the sense that only homotopy equivalences have been removed.

3. The space of sections of a tame function

Let XX be a compact cobordism of manifolds with boundary (Definition 2.1) and f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] a tame function (Definition 2.2).

Definition 3.1.

A section of ff is a continuous map δ:[s,s+]X\delta:[s_{-},s_{+}]\rightarrow X satisfying fδ=id[s,s+]f\circ\delta=id_{[s_{-},s_{+}]}, i.e., δ(t)f1(t)\delta(t)\in f^{-1}(t) for all t[s,s+]t\in[s_{-},s_{+}]. The space of sections of ff is the space of continuous sections

Γf:={δ:[s,s+]X|fδ=id[s,s+]}\Gamma f:=\{\delta:[s_{-},s_{+}]\rightarrow X\,\,|\,\,f\circ\delta=id_{[s_{-},s_{+}]}\}

with the compact-open topology.

We are interested in computing the connected components π0(Γf)\pi_{0}(\Gamma f) in terms of the homotopy groups of pre-images f1(U)f^{-1}(U) of open sets U[s,s+]U\subset[s_{-},s_{+}]. In practice, we perform constructions with regular fibers f1(t)f^{-1}(t) and their inclusions into cobordisms between them. The formalism we use is the zigzag discretization described in §2.2, which is equivalent to working with preimages of open sets by Lemma 2.7.

The main results of this section are the computation of π0(Γf)\pi_{0}(\Gamma f) in Theorem 3.14 and a similar Theorem 3.15 when XX is a compact manifold with boundary and f:XS1f:X\rightarrow S^{1} is submersive with isolated critical points on the boundary with distinct critical values. The proof is the series of arguments in §3.2. It relies on the section splicing and collapse constructions developed in §3.1.

3.1. Section splicing and collapse

We establish several constructions on homotopy classes of sections and their properties.

Consider a section δ:[s,s+]X\delta:[s_{-},s_{+}]\rightarrow X of ff and a path γ:[0,1]X\gamma:[0,1]\rightarrow X_{-} satisfying γ(1)=δ(s)\gamma(1)=\delta(s_{-}). The idea is to splice γ\gamma into δ\delta in a trivial cobordism near ss_{-}. Precisely, choose ϵ>0\epsilon>0 small enough so that the interval [s,s+ϵ][s_{-},s_{-}+\epsilon] does not contain any critical values. Then, by Lemma 2.7, there is a fiberwise diffeomorphism

φ:f1([s,s+ϵ])X×[s,s+ϵ]\varphi:f^{-1}([s_{-},s_{-}+\epsilon])\cong X_{-}\times[s_{-},s_{-}+\epsilon]

that restricts to the identity XX×{s}.X_{-}\rightarrow X_{-}\times\{s_{-}\}. Here fiberwise means that

f=π[s,s+ϵ]φf=\pi_{[s_{-},s_{-}+\epsilon]}\circ\varphi

on f1([s,s+ϵ])Xf^{-1}([s_{-},s_{-}+\epsilon])\subset X, where π[s,s+ϵ]:X×[s,s+ϵ][s,s+ϵ]\pi_{[s_{-},s_{-}+\epsilon]}:X_{-}\times[s_{-},s_{-}+\epsilon]\rightarrow[s_{-},s_{-}+\epsilon] is the projection. There is also a projection π:X×[s,s+ϵ]X\pi:X_{-}\times[s_{-},s_{-}+\epsilon]\rightarrow X_{-}.

Definition 3.2.

Given a section δ:[s,s+]X\delta:[s_{-},s_{+}]\rightarrow X of ff and a path γ:[0,1]X\gamma:[0,1]\rightarrow X_{-} satisfying γ(1)=δ(s)\gamma(1)=\delta(s_{-}), the left boundary ϵ\epsilon-splicing

(γ#ϵδ):[s,s+]X(\gamma\,\,\#_{-}^{\epsilon}\,\,\delta):[s_{-},s_{+}]\rightarrow X

is the section of ff defined by

(γ#ϵδ)(t)={φ1(γ(2ϵ(ts)),t) if t[s,s+ϵ/2]φ1(π(φ(δ(2tsϵ))),t) if t[s+ϵ/2,s+ϵ]δ(t) if t[s+ϵ,s+].(\gamma\,\,\#_{-}^{\epsilon}\,\,\delta)(t)=\begin{cases}\varphi^{-1}(\gamma(\frac{2}{\epsilon}(t-s_{-})),t)&\text{ if }t\in[s_{-},s_{-}+\epsilon/2]\\ \varphi^{-1}(\pi(\varphi(\delta(2t-s_{-}-\epsilon))),t)&\text{ if }t\in[s_{-}+\epsilon/2,s_{-}+\epsilon]\\ \delta(t)&\text{ if }t\in[s_{-}+\epsilon,s_{+}].\\ \end{cases}

For a path γ:[0,1]X+\gamma:[0,1]\rightarrow X_{+} satisfying γ(0)=δ(s+)\gamma(0)=\delta(s_{+}), there is a right boundary ϵ\epsilon-splicing δ#+ϵγ\delta\,\,\#_{+}^{\epsilon}\,\,\gamma defined by gluing γ\gamma into a regular cobordism near s+s_{+} using the symmetric formula.

Given a regular value s(s,s+)s_{*}\in(s_{-},s_{+}) and a path γ:[0,1]Xs=f1(s)\gamma:[0,1]\rightarrow X_{s_{*}}=f^{-1}(s_{*}) satisfying γ(1)=δ(s)\gamma(1)=\delta(s_{*}), the interior ϵ\epsilon-splicing of the path γ\gamma into δ\delta at ss_{*} is the section of ff given by

ϵ(δ,γ,s)\displaystyle{\mathcal{I}}^{\epsilon}(\delta,\gamma,s_{*}) :=(δ|[s,s]#+γ¯,γ#δ|[s,s+])\displaystyle:=(\delta|_{[s_{-},s_{*}]}\,\,\#_{+}\,\,\overline{\gamma},\,\,\gamma\,\,\#_{-}\,\,\delta|_{[s_{*},s_{+}]})
={(δ|[s,s]#+γ¯)(t) if t[s,s](γ#δ|[s,s+])(t) if t[s,s+],\displaystyle=\begin{cases}(\delta|_{[s_{-},s_{*}]}\,\,\#_{+}\,\,\overline{\gamma})(t)&\text{ if }t\in[s_{-},s_{*}]\\ (\gamma\,\,\#_{-}\,\,\delta|_{[s_{*},s_{+}]})(t)&\text{ if }t\in[s_{*},s_{+}],\end{cases}

where

γ¯(t):=γ(1t)\overline{\gamma}(t):=\gamma(1-t)

is the reverse path, δ|[s,s]:[s,s]f1([s,s])\delta|_{[s_{-},s_{*}]}:[s_{-},s_{*}]\rightarrow f^{-1}([s_{-},s_{*}]) is viewed as a section of f1([s,s])[s,s]f^{-1}([s_{-},s_{*}])\rightarrow[s_{-},s_{*}], and similarly for δ|[s,s+]\delta|_{[s_{*},s_{+}]}.

The left boundary ϵ\epsilon-splicing (γ#ϵδ):[s,s+]X(\gamma\,\,\#_{-}^{\epsilon}\,\,\delta):[s_{-},s_{+}]\rightarrow X is a section of ff with endpoints

(γ#ϵδ)(s)=γ(0) and (γ#ϵδ)(s+)=δ(s+).(\gamma\,\,\#_{-}^{\epsilon}\,\,\delta)(s_{-})=\gamma(0)\text{ and }(\gamma\,\,\#_{-}^{\epsilon}\,\,\delta)(s_{+})=\delta(s_{+}).

Similarly, the right boundary splicing has endpoints

(δ#+ϵγ)(s)=δ(s) and (δ#+ϵγ)(s+)=γ(1).(\delta\,\,\#_{+}^{\epsilon}\,\,\gamma)(s_{-})=\delta(s_{-})\text{ and }(\delta\,\,\#_{+}^{\epsilon}\,\,\gamma)(s_{+})=\gamma(1).
Lemma 3.3.

Consider a section δΓf\delta\in\Gamma f of f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] and the constant paths γδ(s)\gamma_{-}\equiv\delta(s_{-}) and γ+δ(s+)\gamma_{+}\equiv\delta(s_{+}). Then the splicings γ#ϵδ\gamma_{-}\,\,\#_{-}^{\epsilon}\,\,\delta and δ#+ϵγ+\delta\,\,\#_{+}^{\epsilon}\gamma_{+} are fiberwise homotopic to δ\delta.

Proof.

A fiberwise homotopy h:[0,1]×[s,s+]Xh:[0,1]\times[s_{-},s_{+}]\rightarrow X from δ\delta to γ#ϵδ\gamma_{-}\,\,\#_{-}^{\epsilon}\,\,\delta is given by

h(r,t):={φ1(δ(s),t) if t[s,s+rϵ/2]φ1(π(φ(δ(2tsrϵ))),t) if t[s+rϵ/2,s+rϵ]δ(t) if t[s+rϵ,s+].\displaystyle h(r,t):=\begin{cases}\varphi^{-1}(\delta(s_{-}),t)&\text{ if }t\in[s_{-},s_{-}+r\epsilon/2]\\ \varphi^{-1}(\pi(\varphi(\delta(2t-s_{-}-r\epsilon))),t)&\text{ if }t\in[s_{-}+r\epsilon/2,s_{-}+r\epsilon]\\ \delta(t)&\text{ if }t\in[s_{-}+r\epsilon,s_{+}].\\ \end{cases}

As a first application of splicings, we use them in the proof of the following lemma to modify a section to have specific behavior near the endpoints.

Lemma 3.4.

Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function. Assume that either all critical points of f|Bf|_{B} are type D, or all critical points are type N. Let δX\delta_{-}\in X_{-} and δ+X+\delta_{+}\in X_{+} such that δ\delta_{-} and δ+\delta_{+} are in the same connected component of XX. Then there exists a continuous section δ:[s,s+]X\delta:[s_{-},s_{+}]\rightarrow X of ff such that δ(s)=δ\delta(s_{-})=\delta_{-} and δ(s+)=δ+\delta(s_{+})=\delta_{+}.

Proof.

Assume that all critical points of f|Bf|_{B} are type D; the argument for type N critical points is symmetric.

Let H:X×[0,s+s]XH:X\times[0,s_{+}-s_{-}]\rightarrow X be the deformation retraction from Lemma 2.6 satisfying

f(H(x,t))=max{s,f(x)t}.f(H(x,t))=\max\{s_{-},\,\,f(x)-t\}.

Define the section

δ~:[s,s+]\displaystyle\tilde{\delta}:[s_{-},s_{+}] X\displaystyle\rightarrow X
t\displaystyle t H(δ+,s+t).\displaystyle\mapsto H(\delta_{+},s_{+}-t).

of ff. It satisfies δ~(s+)=H(δ+,0)=δ+\tilde{\delta}(s_{+})=H(\delta_{+},0)=\delta_{+} and δ~(s)X=f1(s)\tilde{\delta}(s_{-})\in X_{-}=f^{-1}(s_{-}).

We now modify δ~\tilde{\delta} near ss_{-} using a left boundary splicing to obtain a section δ:[s,s+]X\delta:[s_{-},s_{+}]\rightarrow X that also satisfies δ(s)=δ\delta(s_{-})=\delta_{-}. Indeed, since δ~(t)\tilde{\delta}(t) is in the same component of XX as δ+=δ~(s+)\delta_{+}=\tilde{\delta}(s_{+}) for all t[s,s+]t\in[s_{-},s_{+}], which is in the same component of XX as δ\delta_{-} by hypothesis, it follows that δ~(s)\tilde{\delta}(s_{-}) is in the same component of XX as δ\delta_{-}. Taking a path in XX from δ\delta_{-} to δ~(s)\tilde{\delta}(s_{-}) and deformation retracting it with HH into XX_{-} produces a path γ:[0,1]X\gamma:[0,1]\rightarrow X_{-} such that γ(0)=δ\gamma(0)=\delta_{-} and γ(1)=δ~(s)\gamma(1)=\tilde{\delta}(s_{-}). Then the left boundary ϵ\epsilon-splicing (Definition 3.2)

δ:=(γ#ϵδ~):[s,s+]X\delta:=(\gamma\,\,\#_{-}^{\epsilon}\,\,\tilde{\delta}):[s_{-},s_{+}]\rightarrow X

for any sufficiently small ϵ>0\epsilon>0 is a section of ff satisfying δ(s)=γ(0)=δ\delta(s_{-})=\gamma(0)=\delta_{-} and δ(s+)=δ~(s+)=δ+\delta(s_{+})=\tilde{\delta}(s_{+})=\delta_{+}, as required. ∎

Splicings are also useful for converting fiberwise homotopies of sections with free endpoints to fiberwise homotopies with fixed endpoints.

Lemma 3.5.

Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function with sections δ,δ^Γf\delta,\hat{\delta}\in\Gamma f. Suppose there is a fiberwise homotopy h:[0,1]×[s,s+]Xh:[0,1]\times[s_{-},s_{+}]\rightarrow X from δ^\hat{\delta} to δ\delta, meaning that f(h(r,t))=tf(h(r,t))=t, δ^(t)=h(0,t)\hat{\delta}(t)=h(0,t), and δ(t)=h(1,t)\delta(t)=h(1,t) for all r,tr,t. Consider the paths γ±:=h(,s±):[0,1]X±\gamma_{\pm}:=h(-,s_{\pm}):[0,1]\rightarrow X_{\pm}. Then there is a fiberwise homotopy h~:[0,1]×[s,s+]X\tilde{h}:[0,1]\times[s_{-},s_{+}]\rightarrow X from γ#ϵδ#+ϵγ¯+\gamma_{-}\,\,\#_{-}^{\epsilon}\,\,\delta\,\,\#_{+}^{\epsilon}\,\,\overline{\gamma}_{+} to δ^\hat{\delta} that is endpoint preserving, i.e. h~(r,s±)=δ^(s±)\tilde{h}(r,s_{\pm})=\hat{\delta}(s_{\pm}) for all r[0,1]r\in[0,1].

Proof.

For r[0,1]r\in[0,1], consider the paths γ±r(t):=γ±(rt).\gamma_{\pm}^{r}(t):=\gamma_{\pm}(rt). Let c±(t):=δ^(s±)c_{\pm}(t):=\hat{\delta}(s_{\pm}) be the constant paths at the endpoints of δ^\hat{\delta}. Then h~(r,):=γr#ϵh(r,)#+ϵγ+r¯\tilde{h}(r,\cdot):=\gamma^{r}_{-}\,\,\#_{-}^{\epsilon}\,\,h(r,\cdot)\,\,\#_{+}^{\epsilon}\,\,\overline{\gamma^{r}_{+}} is a fiberwise homotopy from h~(0,)=c#ϵδ^#+ϵc+\tilde{h}(0,\cdot)=c_{-}\,\,\#_{-}^{\epsilon}\,\,\hat{\delta}\,\,\#_{+}^{\epsilon}\,\,c_{+} to h~(1,)=γ#ϵδ#+ϵγ¯+\tilde{h}(1,\cdot)=\gamma_{-}\,\,\#_{-}^{\epsilon}\,\,\delta\,\,\#_{+}^{\epsilon}\,\,\overline{\gamma}_{+} and is endpoint preserving. Finally, the section c#ϵδ^#+ϵc+c_{-}\,\,\#_{-}^{\epsilon}\,\,\hat{\delta}\,\,\#_{+}^{\epsilon}\,\,c_{+} is fiberwise homotopic to δ^\hat{\delta} by Lemma 3.3, and the homotopy preserves endpoints. ∎

The next lemma shows that boundary splicing and interior splicing do not affect the fiberwise homotopy class of a section with no endpoint constraints.

Lemma 3.6.

Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function with a section δΓf\delta\in\Gamma f, paths γ±:[0,1]X±\gamma_{\pm}:[0,1]\rightarrow X_{\pm} such that γ(1)=δ(s)\gamma_{-}(1)=\delta(s_{-}) and γ+(0)=δ(s+)\gamma_{+}(0)=\delta(s_{+}), and a path γ:[0,1]f1(s)\gamma:[0,1]\rightarrow f^{-1}(s_{*}) for some regular value ss_{*} of f|Bf|_{B} such that γ(1)=δ(s)\gamma(1)=\delta(s_{*}). Then the left boundary splicing γ#ϵδ\gamma_{-}\,\,\#_{-}^{\epsilon}\,\,\delta, the right boundary splicing δ#+ϵγ+\delta\,\,\#_{+}^{\epsilon}\,\,\gamma_{+}, and the interior splicing ϵ(δ,γ,s){\mathcal{I}}^{\epsilon}(\delta,\gamma,s_{*}) are all fiberwise homotopic with free endpoints to δ\delta. That is, all of these sections are in the same path component of Γf\Gamma f as δ\delta.

Proof.

Since the endpoints are allowed to move on X±=f1(s±)X_{\pm}=f^{-1}(s_{\pm}) during the homotopy, we can ‘unwind’ the path γ±\gamma_{\pm} spliced onto the left or the right in a trivial cobordism near s±s_{\pm}. The interior splicing ϵ(δ,γ,s){\mathcal{I}}^{\epsilon}(\delta,\gamma,s_{*}) inserts the graph of the path γ\gamma next to the graph of γ¯\overline{\gamma} in a trivial cobordism near f1(s)f^{-1}(s_{*}). The concatenation of these paths in f1(s)f^{-1}(s_{*}) is a loop based at δ(s)\delta(s_{*}) and is contractible relative the basepoint. The graph of this contraction provides a fiberwise homotopy from the interior splicing to δ\delta. ∎

We now introduce some notation. Given points xXx_{-}\in X_{-} and x+X+x_{+}\in X_{+}, define the space of sections with fixed endpoints

Γfxx+:={δΓf|δ(s)=x and δ(s+)=x+}.\Gamma f_{x_{-}}^{x^{+}}:=\{\delta\in\Gamma f\,\,|\,\,\delta(s_{-})=x_{-}\text{ and }\delta(s_{+})=x_{+}\}.

Given a space AA and points x,xAx,x^{\prime}\in A, the path space with fixed endpoints is denoted

𝒫(A)xx:={γ:[0,1]A|γ(0)=x and γ(1)=x}.{\mathcal{P}}(A)^{x}_{x^{\prime}}:=\{\gamma:[0,1]\rightarrow A\,\,|\,\,\gamma(0)=x^{\prime}\text{ and }\gamma(1)=x\}.

The based loop space of AA at xx is denoted

Ω(A)x=𝒫(A)xx.\Omega(A)_{x}={\mathcal{P}}(A)_{x}^{x}.

The fiberwise homotopy class of a left boundary splicing γ#ϵδ\gamma\,\,\#_{-}^{\epsilon}\,\,\delta relative its endpoints [γ#ϵδ]π0(Γfγ(0)δ(s+))[\gamma\,\,\#_{-}^{\epsilon}\,\,\delta]\in\pi_{0}(\Gamma f_{\gamma(0)}^{\delta(s_{+})}) does not depend on the homotopy class of γ\gamma relative its endpoints [γ]π0(𝒫(X)γ(0)γ(1))[\gamma]\in\pi_{0}({\mathcal{P}}(X_{-})_{\gamma(0)}^{\gamma(1)}), the fiberwise homotopy class of δ\delta relative its endpoints [δ]π0(Γfδ(s)δ(s+))[\delta]\in\pi_{0}(\Gamma f_{\delta(s_{-})}^{\delta(s_{+})}), or ϵ>0\epsilon>0. In particular, for any x,xXx_{-}^{\prime},x_{-}\in X_{-} and x+X+x_{+}\in X_{+}, left boundary ϵ\epsilon-splicing descends to a well-defined map

#:π0(𝒫(X)xx)×π0(Γfxx+)\displaystyle\#_{-}:\pi_{0}({\mathcal{P}}(X_{-})_{x_{-}^{\prime}}^{x_{-}})\times\pi_{0}(\Gamma f_{x_{-}}^{x_{+}}) π0(Γfxx+)\displaystyle\rightarrow\pi_{0}(\Gamma f_{x_{-}^{\prime}}^{x_{+}})
([γ],[δ])\displaystyle([\gamma],[\delta]) [γ#ϵδ].\displaystyle\mapsto[\gamma\,\,\#_{-}^{\epsilon}\,\,\delta].

In the case x=xx^{\prime}_{-}=x_{-} the path space 𝒫(X)xx{\mathcal{P}}(X_{-})_{x_{-}^{\prime}}^{x_{-}} is equal to the based loop space Ω(X)x\Omega(X_{-})_{x_{-}} and moreover π0(𝒫(X)xx)=π1(X,x)\pi_{0}({\mathcal{P}}(X_{-})_{x_{-}^{\prime}}^{x_{-}})=\pi_{1}(X_{-},x_{-}). The map

#:π1(X,x)×π0(Γfxx+)\displaystyle\#_{-}:\pi_{1}(X_{-},x_{-})\times\pi_{0}(\Gamma f_{x_{-}}^{x_{+}}) π0(Γfxx+)\displaystyle\rightarrow\pi_{0}(\Gamma f_{x_{-}}^{x_{+}})

is a left group action of π1(X,x)\pi_{1}(X_{-},x_{-}) on the set π0(Γfxx+).\pi_{0}(\Gamma f_{x_{-}}^{x_{+}}).

Similarly, given xXx_{-}\in X_{-} and x+,x+X+x_{+},x_{+}^{\prime}\in X_{+}, right boundary ϵ\epsilon-splicing descends to a well-defined map

#+:π0(Γfxx+)×π0(𝒫(X+)x+x+)\displaystyle\#_{+}:\pi_{0}(\Gamma f_{x_{-}}^{x_{+}})\times\pi_{0}({\mathcal{P}}(X_{+})_{x_{+}}^{x_{+}^{\prime}}) π0(Γfxx+)\displaystyle\rightarrow\pi_{0}(\Gamma f_{x_{-}}^{x_{+}^{\prime}})
([δ],[γ])\displaystyle([\delta],[\gamma]) [δ#+ϵγ].\displaystyle\mapsto[\delta\,\,\#_{+}^{\epsilon}\,\,\gamma].

In the case x+=x+x^{\prime}_{+}=x_{+}, the map

#+:π0(Γfxx+)×π1(X+,x+)\displaystyle\#_{+}:\pi_{0}(\Gamma f_{x_{-}}^{x_{+}})\times\pi_{1}(X_{+},x_{+}) π0(Γfxx+)\displaystyle\rightarrow\pi_{0}(\Gamma f_{x_{-}}^{x_{+}})

is a right group action of π1(X+,x+)\pi_{1}(X_{+},x_{+}) on the set π0(Γfxx+).\pi_{0}(\Gamma f_{x_{-}}^{x_{+}}).

We now define various versions of a collapse map and establish their properties. Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function with only type DD critical points or, symmetrically, with only type NN critical points. For the rest of this section, the symbol ±\pm means - if type DD and ++ if type NN, and symmetrically for the symbol \mp.

Fix a basepoint section 𝔟Γf\mathfrak{b}\in\Gamma f. Let H:X×[0,s+s]XH:X\times[0,s_{+}-s_{-}]\rightarrow X be a deformation retraction along 𝔟\mathfrak{b} and onto X±X_{\pm}, as provided by Lemma 2.6. Then define the cobordism collapse map

(5) C:=H(,s+s):XX±.C:=H(-,s_{+}-s_{-}):X\rightarrow X_{\pm}.

The map CC is a left inverse of the inclusion X±XX_{\pm}\rightarrow X. Since H(𝔟(t),s+s)=𝔟(s±)H(\mathfrak{b}(t),s_{+}-s_{-})=\mathfrak{b}(s_{\pm}) for all t[s,s+]t\in[s_{-},s_{+}] by Lemma 2.6, we have

C(𝔟(t))=𝔟(s±) for all t[s,s+],C(\mathfrak{b}(t))=\mathfrak{b}(s_{\pm})\text{ for all }t\in[s_{-},s_{+}],

and so CC is a map of pairs

C:(X,𝔟([s,s+]))(X±,𝔟(s±)).C:(X,\mathfrak{b}([s_{-},s_{+}]))\rightarrow(X_{\pm},\mathfrak{b}(s_{\pm})).

This map induces the section collapse map on components

(6) ΓC:π0(Γf𝔟(s)𝔟(s+))\displaystyle\Gamma C_{*}:\pi_{0}(\Gamma f_{\mathfrak{b}(s_{-})}^{\mathfrak{b}(s_{+})}) π1(X±,𝔟(s±))\displaystyle\rightarrow\pi_{1}(X_{\pm},\mathfrak{b}(s_{\pm}))
[δ]\displaystyle[\delta] [Cδ].\displaystyle\mapsto[C\circ\delta].

We prove in Lemma 3.8 that ΓC\Gamma C_{*} is a bijection. There is a similarly defined loop collapse map

C:π1(X,𝔟(s))\displaystyle C_{*}:\pi_{1}(X_{\mp},\mathfrak{b}(s_{\mp})) π1(X±,𝔟(s±))\displaystyle\rightarrow\pi_{1}(X_{\pm},\mathfrak{b}(s_{\pm}))
[γ]\displaystyle[\gamma] [Cγ].\displaystyle\mapsto[C\circ\gamma].

Section collapse ΓC\Gamma C_{*} and section splicing #±\#_{\pm} are related as follows.

Lemma 3.7.

Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function with only type DD critical points and 𝔟Γf\mathfrak{b}\in\Gamma f. Let C:XXC:X\rightarrow X_{-} be an associated collapse map along 𝔟\mathfrak{b}.

Consider a section δΓf𝔟(s)𝔟(s+)\delta\in\Gamma f_{\mathfrak{b}(s_{-})}^{\mathfrak{b}(s_{+})}. Then, in π0(Γf𝔟(s)𝔟(s+))\pi_{0}(\Gamma f_{\mathfrak{b}(s_{-})}^{\mathfrak{b}(s_{+})}), we have

ΓC([δ])#[𝔟]=[δ].\Gamma C_{*}([\delta])\,\,\#_{-}\,\,[\mathfrak{b}]=[\delta].

Moreover, consider loops γπ1(X,𝔟(s))\gamma_{-}\in\pi_{1}(X_{-},\mathfrak{b}(s_{-})) and γ+π1(X+,𝔟(s+))\gamma_{+}\in\pi_{1}(X_{+},\mathfrak{b}(s_{+})). Then, in π1(X,𝔟(s))\pi_{1}(X_{-},\mathfrak{b}(s_{-})) with group operation denoted \cdot, it holds that

ΓC(γ#δ#+γ+)=γΓC(δ)C(γ+).\Gamma C_{*}(\gamma_{-}\,\,\#_{-}\,\,\delta\,\,\#_{+}\,\,\gamma_{+})=\gamma_{-}\cdot\Gamma C_{*}(\delta)\cdot C_{*}(\gamma_{+}).

In particular,

ΓC(γ#δ)=γΓC(δ)\Gamma C_{*}(\gamma_{-}\,\,\#_{-}\,\,\delta)=\gamma_{-}\cdot\Gamma C_{*}(\delta)

and

ΓC(δ#+γ+)=ΓC(δ)C(γ+).\Gamma C_{*}(\delta\,\,\#_{+}\,\,\gamma_{+})=\Gamma C_{*}(\delta)\cdot C_{*}(\gamma_{+}).

The symmetric statement holds if ff has only type NN critical points.

Proof.

To prove ΓC([δ])#[𝔟]=[δ],\Gamma C_{*}([\delta])\,\,\#_{-}\,\,[\mathfrak{b}]=[\delta], we must show that for small ϵ>0\epsilon>0 there is a fiberwise homotopy from δ\delta to the section (Cδ)#ϵ𝔟(C\circ\delta)\,\,\#_{-}^{\epsilon}\,\,\mathfrak{b}. The section (Cδ)#ϵ𝔟(C\circ\delta)\,\,\#_{-}^{\epsilon}\,\,\mathfrak{b} is given by flowing δ\delta to the left until it lies in a trivial cobordism near XX_{-} with its right endpoint lying on 𝔟\mathfrak{b} and then attaching this onto the left of 𝔟\mathfrak{b}. The homotopy is constructed using this flow. We now write this out in precise formulas.

For simplification of the formulas, assume without loss of generality that s=0s_{-}=0 and s+=1s_{+}=1. By definition, we have

(Cδ)(t)=H(δ(t),1) for t[0,1](C\circ\delta)(t)=H(\delta(t),1)\text{ for }t\in[0,1]

and

((Cδ)#ϵ𝔟)(t)={φ1(H(δ(2ϵt),1),t) if t[0,ϵ/2]φ1(π(φ(𝔟(2tϵ))),t) if t[ϵ/2,ϵ]𝔟(t) if t[ϵ,1].((C\circ\delta)\,\,\#_{-}^{\epsilon}\,\,\mathfrak{b})(t)=\begin{cases}\varphi^{-1}(H(\delta(\frac{2}{\epsilon}t),1),t)&\text{ if }t\in[0,\epsilon/2]\\ \varphi^{-1}(\pi(\varphi(\mathfrak{b}(2t-\epsilon))),t)&\text{ if }t\in[\epsilon/2,\epsilon]\\ \mathfrak{b}(t)&\text{ if }t\in[\epsilon,1].\end{cases}

The claimed homotopy is then given by

(r,t){φ1(π(φ(H(δ(2trϵ(t+r(1t))),r))),t) if t[0,rϵ2]φ1(π(φ(H(δ(t+r(1t)),r(1(2trϵ))))),t) if t[rϵ2,rϵ]H(δ(t+r(1t)),r(1t)) if t[rϵ,1].(r,t)\mapsto\begin{cases}\varphi^{-1}(\pi(\varphi(H(\delta(\frac{2t}{r\epsilon}(t+r(1-t))),r))),t)&\text{ if }t\in[0,\frac{r\epsilon}{2}]\\ \varphi^{-1}(\pi(\varphi(H(\delta(t+r(1-t)),r(1-(2t-r\epsilon))))),t)&\text{ if }t\in[\frac{r\epsilon}{2},r\epsilon]\\ H(\delta(t+r(1-t)),r(1-t))&\text{ if }t\in[r\epsilon,1].\end{cases}

for r[0,1]r\in[0,1]. Note that the rr in the denominator in the case t[0,rϵ2]t\in[0,\frac{r\epsilon}{2}] is not a continuity issue because we have

02trϵ(t+r(1t))=2trϵ(t(1r)+r)rϵ2(1r)+r0\leq\frac{2t}{r\epsilon}(t+r(1-t))=\frac{2t}{r\epsilon}(t(1-r)+r)\leq\frac{r\epsilon}{2}(1-r)+r

and hence the limit as r0r\rightarrow 0 is equal to zero. This completes the proof of the claim ΓC([δ])#[𝔟]=[δ]\Gamma C_{*}([\delta])\,\,\#_{-}\,\,[\mathfrak{b}]=[\delta].

The next claim ΓC(γ#δ#+γ+)=γΓC(δ)C(γ+)\Gamma C_{*}(\gamma_{-}\,\,\#_{-}\,\,\delta\,\,\#_{+}\,\,\gamma_{+})=\gamma_{-}\cdot\Gamma C_{*}(\delta)\cdot C_{*}(\gamma_{+}) is seen as follows. The left side is obtained by flowing γ#δ#+γ+\gamma_{-}\,\,\#_{-}\,\,\delta\,\,\#_{+}\,\,\gamma_{+} to the left until it lies in XX_{-}. The result of this is the loop γ\gamma_{-}, followed by the section δ\delta pushed into XX_{-}, which is exactly ΓC(δ)\Gamma C_{*}(\delta), followed by the loop γ+\gamma_{+} pushed from X+X_{+} into XX_{-} via the flow, which is C(γ+)C_{*}(\gamma_{+}). Appending one path after another is the composition \cdot in π1(X,𝔟(s))\pi_{1}(X_{-},\mathfrak{b}(s_{-})) on the right hand side of the claimed formula. ∎

An important consequence of Lemma 3.7 is that the section collapse map ΓC\Gamma C_{*} is bijective.

Lemma 3.8.

Let f:X[s,s+]f:X\rightarrow[s_{-},s_{+}] be a tame function with only type DD critical points or only type NN critical points, and let 𝔟Γf\mathfrak{b}\in\Gamma f. Then the section collapse map ΓC\Gamma C_{*} defined in (6) is bijective.

Proof.

Assume all critical points are type DD. For surjectivity, let γπ1(X,𝔟(s))\gamma\in\pi_{1}(X_{-},\mathfrak{b}(s_{-})) and consider the left boundary splicing δ:=γ#𝔟.\delta:=\gamma\,\,\#_{-}\,\,\mathfrak{b}. Then by Lemma 3.7 we have ΓC(δ)=γΓC(𝔟)=γ1=γ\Gamma C_{*}(\delta)=\gamma\cdot\Gamma C_{*}(\mathfrak{b})=\gamma\cdot 1=\gamma. For injectivity, let δ,δ^π0(Γf𝔟(s)𝔟(s+))\delta,\hat{\delta}\in\pi_{0}(\Gamma f_{\mathfrak{b}(s_{-})}^{\mathfrak{b}(s_{+})}) and assume ΓC(δ)=ΓC(δ^).\Gamma C_{*}(\delta)=\Gamma C_{*}(\hat{\delta}). Then by Lemma 3.7 we have δ=ΓC(δ)#[𝔟]=ΓC(δ^)#[𝔟]=δ^.\delta=\Gamma C_{*}(\delta)\,\,\#_{-}\,\,[\mathfrak{b}]=\Gamma C_{*}(\hat{\delta})\,\,\#_{-}\,\,[\mathfrak{b}]=\hat{\delta}.

3.2. Connected components

The connected components π0(Γf)\pi_{0}(\Gamma f) of the space of sections Γf\Gamma f of a tame function f:XI=[0,1]f:X\rightarrow I=[0,1] can be computed from π0\pi_{0} and π1\pi_{1} of the regular fibers of ff and maps between them coming from the regular cobordisms between them. This is the main theorem (Theorem 3.14) proved in this section. There is a similar statement for maps XS1X\rightarrow S^{1} given in Theorem 3.15.

Recall the diagram ZXZX of regular fibers of ff from (4). Applying π0\pi_{0} produces a diagram of sets

π0(ZX)=(π0(X0)π0(X1)π0(Xn)).\pi_{0}(ZX)=\bigg{(}\pi_{0}(X_{0})\leftrightarrow\pi_{0}(X_{1})\leftrightarrow\cdots\leftrightarrow\pi_{0}(X_{n})\bigg{)}.

A central object in this analysis is the inverse limit limπ0(ZX)\varprojlim\pi_{0}(ZX). Recall that an element Ψlimπ0(ZX)\Psi\in\varprojlim\pi_{0}(ZX) is a collection

Ψsi\displaystyle\Psi_{s_{i}} π0(Xi) for 0in,\displaystyle\in\pi_{0}(X_{i})\text{ for }0\leq i\leq n,

such that, for all ii, it holds that βi,i+1(Ψs(i))=Ψs+(i){\beta_{i,i+1}}_{*}(\Psi_{s_{\partial^{-}(i)}})=\Psi_{s_{\partial^{+}(i)}} where βi,i+1:π0(X(i))π0(X+(i)){\beta_{i,i+1}}_{*}:\pi_{0}(X_{\partial^{-}(i)})\rightarrow\pi_{0}(X_{\partial^{+}(i)}) is the map on π0\pi_{0} induced by the map βi,i+1\beta_{i,i+1} defined in (3).

The connected components π0(Γf)\pi_{0}(\Gamma f) of the space of sections Γf\Gamma f are related to the inverse limit limπ0(ZX)\varprojlim\pi_{0}(ZX) via the map

Π0:π0(Γf)\displaystyle\Pi_{0}:\pi_{0}(\Gamma f) limπ0(ZX)\displaystyle\rightarrow\varprojlim\pi_{0}(ZX)
[δ]\displaystyle[\delta] {Ψsi:=[δ(si)]π0(Xi) for i=0,,n}.\displaystyle\mapsto\{\Psi_{s_{i}}:=[\delta(s_{i})]\in\pi_{0}(X_{i})\text{ for }i=0,\ldots,n\}.

Note that π0(Γf)\pi_{0}(\Gamma f) is the set of fiber-preserving homotopy classes of sections of ff.

The map Π0\Pi_{0} is surjective, as we prove in Proposition 3.9. The idea of the proof is as follows. We must lift a given Ψlimπ0(ZX)\Psi\in\varprojlim\pi_{0}(ZX) through Π0\Pi_{0} to a section IXI\rightarrow X of ff. For each cobordism Xii+1=f1([si,si+1])X_{i}^{i+1}=f^{-1}([s_{i},s_{i+1}]) of manifolds with boundary Xi=f1(si)X_{i}=f^{-1}(s_{i}) and Xi+1=f1(si+1)X_{i+1}=f^{-1}(s_{i+1}), Lemma 3.4 provides a lift over the interval [si,si+1][s_{i},s_{i+1}]. These lifts agree at the regular values sis_{i}, hence they fit together to form the desired lift of Ψ\Psi.

Proposition 3.9.

The map Π0:π0(Γf)limπ0(ZX)\Pi_{0}:\pi_{0}(\Gamma f)\rightarrow\varprojlim\pi_{0}(ZX) is surjective.

Proof.

Let Ψlimπ0(ZX)\Psi\in\varprojlim\pi_{0}(ZX) given by Ψsiπ0(Xi) for 0in\Psi_{s_{i}}\in\pi_{0}(X_{i})\text{ for }0\leq i\leq n. To prove the lemma, we must lift Ψ\Psi through Π0\Pi_{0} to a continuous section δ:[0,1]X\delta:[0,1]\rightarrow X satisfying

(7) [δ(si)]=Ψsi[\delta(s_{i})]=\Psi_{s_{i}}

for all 0in0\leq i\leq n.

For each 0in0\leq i\leq n, choose a lift of Ψsi\Psi_{s_{i}}, i.e. a point

δiΨsi.\delta_{i}\in\Psi_{s_{i}}.

For each 0in10\leq i\leq n-1, the restriction

f|Xii+1:Xii+1[si,si+1]f|_{X_{i}^{i+1}}:X_{i}^{i+1}\rightarrow[s_{i},s_{i+1}]

is a tame function on the compact cobordism Xii+1X_{i}^{i+1} between the manifolds with boundary XiX_{i} and Xi+1X_{i+1}. We claim that the hypotheses of Lemma 3.4 hold, providing a continuous section

δi,i+1:[si,si+1]Xii+1\delta_{i,i+1}:[s_{i},s_{i+1}]\rightarrow X_{i}^{i+1}

such that

δi,i+1(si)=δi and δi,i+1(si+1)=δsi+1.\delta_{i,i+1}(s_{i})=\delta_{i}\text{ and }\delta_{i,i+1}(s_{i+1})=\delta_{s_{i+1}}.

Indeed, since [si,si+1][s_{i},s_{i+1}] contains exactly 11 critical value of f|Bf|_{B} by the choice of interleaving regular values sis_{i}, the function f|Xii+1f|_{X_{i}^{i+1}} has exactly 11 critical point on the boundary. Moreover, δi\delta_{i} and δi+1\delta_{i+1} are contained in the same connected component of Xii+1X_{i}^{i+1} since βi,i+1(Ψs(i))=Ψs+(i){\beta_{i,i+1}}_{*}(\Psi_{s_{\partial^{-}(i)}})=\Psi_{s_{\partial^{+}(i)}}.

The endpoints of the sections δi,i+1\delta_{i,i+1} agree, i.e., δi,i+1(si+1)=δi+1,i+2(si+1)\delta_{i,i+1}(s_{i+1})=\delta_{i+1,i+2}(s_{i+1}) for all 0in20\leq i\leq n-2. Hence they fit together to form a continuous section δ:IX\delta:I\rightarrow X of ff which satisfies (7). Hence Π0([δ])=Ψ\Pi_{0}([\delta])=\Psi and the proof is complete. ∎

To calculate π0(Γf)\pi_{0}(\Gamma f), it remains to characterize the fibers of the surjection Π0:π0(Γf)limπ0(ZX)\Pi_{0}:\pi_{0}(\Gamma f)\rightarrow\varprojlim\pi_{0}(ZX).

Let Ψlimπ0(ZX)\Psi\in\varprojlim\pi_{0}(ZX). Choose a fixed ‘basepoint’ section 𝔟Γf\mathfrak{b}\in\Gamma f such that

Π0([𝔟])=Ψ,\Pi_{0}([\mathfrak{b}])=\Psi,

which means that

𝔟(si)Ψsi for 0in.\mathfrak{b}(s_{i})\in\Psi_{s_{i}}\text{ for }0\leq i\leq n.

As an intermediate object in our analysis, we consider the subspace of sections that agree with 𝔟\mathfrak{b} at regular values sis_{i}; precisely, we define

Γf(𝔟):={δΓf|δ(si)=𝔟(si) for all i}.\Gamma f(\mathfrak{b}):=\{\delta\in\Gamma f\,\,|\,\,\delta(s_{i})=\mathfrak{b}(s_{i})\text{ for all }i\}.

This space splits as the Cartesian product of the spaces of sections of the restrictions f|Xii+1:Xii+1[si,si+1]f|_{X_{i}^{i+1}}:X_{i}^{i+1}\rightarrow[s_{i},s_{i+1}] with endpoints agreeing with 𝔟\mathfrak{b}, i.e.

Γf(𝔟)=i=0n1Γf|Xii+1(𝔟|[si,si+1]).\Gamma f(\mathfrak{b})=\prod_{i=0}^{n-1}\Gamma f|_{X_{i}^{i+1}}(\mathfrak{b}|_{[s_{i},s_{i+1}]}).

Notice that, for every δΓf(𝔟)\delta\in\Gamma f(\mathfrak{b}), we have Π0([δ])=Ψ\Pi_{0}([\delta])=\Psi. It follows that the inclusion ι:Γf(𝔟)Γf\iota:\Gamma f(\mathfrak{b})\hookrightarrow\Gamma f induces a map

π0(Γf(𝔟))ιΠ01(Ψ)π0(Γf).\pi_{0}(\Gamma f(\mathfrak{b}))\xrightarrow{\iota_{*}}\Pi_{0}^{-1}(\Psi)\subset\pi_{0}(\Gamma f).
Proposition 3.10.

The map π0(Γf(𝔟))ιΠ01(Ψ)\pi_{0}(\Gamma f(\mathfrak{b}))\xrightarrow{\iota_{*}}\Pi_{0}^{-1}(\Psi) is surjective.

Proof.

Let δΓf\delta\in\Gamma f such that [δ]Π01(Ψ)[\delta]\in\Pi_{0}^{-1}(\Psi). This means that δ(si)Ψsi\delta(s_{i})\in\Psi_{s_{i}} for all i=0,,ni=0,\ldots,n. Since also 𝔟(si)Ψsi\mathfrak{b}(s_{i})\in\Psi_{s_{i}} for all ii, there is a continuous path γi:[0,1]Ψsi\gamma_{i}:[0,1]\rightarrow\Psi_{s_{i}} from γi(0)=𝔟(si)\gamma_{i}(0)=\mathfrak{b}(s_{i}) to γi(1)=δ(si)\gamma_{i}(1)=\delta(s_{i}). The idea is to splice γi\gamma_{i} into δ\delta on the left of sis_{i} and to splice γ¯i():=γi(1)\overline{\gamma}_{i}(\cdot):=\gamma_{i}(1-\cdot) into δ\delta on the right of sis_{i}, producing a new section δ^Γf(𝔟)\hat{\delta}\in\Gamma f(\mathfrak{b}), i.e. δ^(si)=𝔟(si)\hat{\delta}(s_{i})=\mathfrak{b}(s_{i}) for all ii. Moreover, in the space of sections Γf\Gamma f with no point restrictions, this spliced section δ^\hat{\delta} is in the same path component as δ\delta, i.e. δ^\hat{\delta} is fiberwise homotopic to δ\delta by Lemma 3.6. Then ι([δ^])=[δ]\iota_{*}([\hat{\delta}])=[\delta] and the proof is complete.

We precisely construct the spliced section δ^\hat{\delta} and verify the claimed properties. Restricting δ\delta to subintervals produces a section δi:[si,si+1]Xii+1\delta_{i}:[s_{i},s_{i+1}]\rightarrow X_{i}^{i+1} of the tame function f|Xii+1[si,si+1]f|_{X_{i}^{i+1}}\rightarrow[s_{i},s_{i+1}] for i=0,,n1i=0,\ldots,n-1. Form the sections

δ^i:=(γi#ϵδi#+ϵγi+1¯):[si,si+1]Xii+1\hat{\delta}_{i}:=\big{(}\gamma_{i}\,\,\#_{-}^{\epsilon}\,\,\delta_{i}\,\,\#_{+}^{\epsilon}\,\,\overline{\gamma_{i+1}}\big{)}:[s_{i},s_{i+1}]\rightarrow X_{i}^{i+1}

for some ϵ>0\epsilon>0 small enough. Then for all 0in10\leq i\leq n-1 it holds that δ^i(si)=𝔟(si)\hat{\delta}_{i}(s_{i})=\mathfrak{b}(s_{i}) and δ^i(si+1)=𝔟(si+1)\hat{\delta}_{i}(s_{i+1})=\mathfrak{b}(s_{i+1}). In particular, δ^i(si+1)=δ^i+1(si+1)\hat{\delta}_{i}(s_{i+1})=\hat{\delta}_{i+1}(s_{i+1}), and hence the δ^i\hat{\delta}_{i} fit together to form a section δ^:[0,1]X\hat{\delta}:[0,1]\rightarrow X of ff. Moreover, δ^Γf(𝔟)\hat{\delta}\in\Gamma f(\mathfrak{b}) since δ^(si)=𝔟(si)\hat{\delta}(s_{i})=\mathfrak{b}(s_{i}) for all ii.

It remains to show that δ^\hat{\delta} is fiberwise homotopic to δ\delta with no point restrictions. Observe that δ^\hat{\delta} is formed by starting with δ\delta, performing a left boundary splicing with γ0\gamma_{0}, then performing an interior splicing with γi\gamma_{i} for all 1in11\leq i\leq n-1, and finally performing a right boundary splicing with γn\gamma_{n}. Hence the claim follows from repeated application of Lemma 3.6. ∎

Consider the direct product group i=0nπ1(Xi,𝔟(si))\prod_{i=0}^{n}\pi_{1}(X_{i},\mathfrak{b}(s_{i})). There is a group action on the set π0(Γf(𝔟))\pi_{0}(\Gamma f(\mathfrak{b})), denoted by \star and defined in (8),

:i=0nπ1(Xi,𝔟(si))×π0(Γf(𝔟))π0(Γf(𝔟)).\star:\prod_{i=0}^{n}\pi_{1}(X_{i},\mathfrak{b}(s_{i}))\times\pi_{0}(\Gamma f(\mathfrak{b}))\rightarrow\pi_{0}(\Gamma f(\mathfrak{b})).

The orbits of this action are exactly the fibers of ι\iota_{*}, as we prove in Proposition 3.11. Hence, due to Proposition 3.10 above, Π01(Ψ)\Pi_{0}^{-1}(\Psi) is naturally in bijection with the set of orbits.

We proceed to define the claimed group action and establish its properties. The action of τ=(τ0,,τn)i=0nπ1(Xi,𝔟(si))\tau=(\tau_{0},\ldots,\tau_{n})\in\prod_{i=0}^{n}\pi_{1}(X_{i},\mathfrak{b}(s_{i})) on a class [δ]π0(Γf(𝔟))[\delta]\in\pi_{0}(\Gamma f(\mathfrak{b})) is defined by representing δ\delta by its restrictions δi:=δ|[si,si+1]Γf|Xii+1(𝔟|[si,si+1])\delta_{i}:=\delta|_{[s_{i},s_{i+1}]}\in\Gamma f|_{X_{i}^{i+1}}(\mathfrak{b}|_{[s_{i},s_{i+1}]}) for i=0,,n1i=0,\ldots,n-1 and taking the spliced homotopy class

(8) τ[δ]:=(τ0#[δ0]#+τ11,τ1#[δ1]#+τ21,,τn1#[δn1]#+τn1).\tau\star[\delta]:=(\tau_{0}\#_{-}[\delta_{0}]\#_{+}\tau_{1}^{-1},\tau_{1}\#_{-}[\delta_{1}]\#_{+}\tau_{2}^{-1},\ldots,\tau_{n-1}\#_{-}[\delta_{n-1}]\#_{+}\tau_{n}^{-1}).
Proposition 3.11.

The orbits of the group action \star are the fibers of ι\iota_{*}. Precisely, this means that 𝒪π0(Γf(𝔟)){\mathcal{O}}\subset\pi_{0}(\Gamma f(\mathfrak{b})) is an orbit of \star if and only if 𝒪=ι1([δ]){\mathcal{O}}=\iota_{*}^{-1}([\delta]) for some [δ]Π01(Ψ)π0(Γf)[\delta]\in\Pi_{0}^{-1}(\Psi)\subset\pi_{0}(\Gamma f).

In particular, ι\iota_{*} induces a bijection between the set of orbits of \star and Π01(Ψ)\Pi_{0}^{-1}(\Psi).

Proof.

Let δ,δ^Γf(𝔟)\delta,\hat{\delta}\in\Gamma f(\mathfrak{b}). It must be shown that δ\delta is fiberwise homotopic to δ^\hat{\delta} if and only if there exists an element τi=0nπi(Xi,𝔟(si))\tau\in\prod_{i=0}^{n}\pi_{i}(X_{i},\mathfrak{b}(s_{i})) satisfying τ[δ]=[δ^]\tau\star[\delta]=[\hat{\delta}].

By the definition (8) of the action, for any τ\tau and δ\delta we see that a representative section of τ[δ]\tau\star[\delta] is a sequence of a left boundary splicing with τ0\tau_{0}, followed by interior splicing of τi\tau_{i} at the regular values sis_{i} for all 1in11\leq i\leq n-1, followed by a right boundary splicing of τn\tau_{n}. Hence by Lemma 3.6 it is fiberwise homotopic to δ\delta. So, if it is assumed that τ[δ]=[δ^]\tau\star[\delta]=[\hat{\delta}], it follows that δ^\hat{\delta} is fiberwise homotopic to δ\delta. This proves one direction of the if and only if statement.

To prove the other direction, suppose that δ\delta is fiberwise homotopic to δ^\hat{\delta}. For each i=0,,n1i=0,\ldots,n-1, this fiberwise homotopy restricts to a homotopy from the section δi:=δ|[si,si+1]:[si,si+1]Xii+1\delta_{i}:=\delta|_{[s_{i},s_{i+1}]}:[s_{i},s_{i+1}]\rightarrow X_{i}^{i+1} to the section δ^i:=δ^|[si,si+1]\hat{\delta}_{i}:=\hat{\delta}|_{[s_{i},s_{i+1}]}. Then, Lemma 3.5 provides loops τi:[0,1]Xi\tau_{i}:[0,1]\rightarrow X_{i} based at 𝔟(si)\mathfrak{b}(s_{i}) for i=0,,ni=0,\ldots,n and a fiberwise homotopy from τi#ϵδi#+ϵτi+11\tau_{i}\,\,\#_{-}^{\epsilon}\,\,\delta_{i}\,\,\#_{+}^{\epsilon}\,\,{\tau}^{-1}_{i+1} to δ^i\hat{\delta}_{i} that preserves the endpoints at 𝔟(si)\mathfrak{b}(s_{i}) and 𝔟(si+1)\mathfrak{b}(s_{i+1}). Since these homotopies are endpoint preserving, setting τ:=(τ0,,τn)\tau:=(\tau_{0},\ldots,\tau_{n}) they fit together to provide a homotopy from a representative of τ[δ]\tau\star[\delta] to δ^\hat{\delta} within the space Γf(𝔟)\Gamma f(\mathfrak{b}). Hence τ[δ]=[δ^]\tau\star[\delta]=[\hat{\delta}], as required. ∎

In light of Proposition 3.11, to calculate Π01(Ψ)\Pi_{0}^{-1}(\Psi) it suffices to understand the group action \star. The group acting is defined in the regular fibers of ff – it is a product of fundamental groups of fibers – however the set π0(Γf(𝔟))\pi_{0}(\Gamma f(\mathfrak{b})) and the action have not yet been algebraically described in terms of homotopy theoretic information about the fibers. We proceed to do this now. First, we bijectively identify π0(Γf(𝔟))\pi_{0}(\Gamma f(\mathfrak{b})) with a product of fundamental groups of fibers (Proposition 3.12), and then we understand the group action \star in terms of composition of loops in these fundamental groups (Proposition 3.13). The resulting characterization of Π01(Ψ)\Pi_{0}^{-1}(\Psi) is summarized in Theorem 3.14.

Every section δΓf(𝔟)\delta\in\Gamma f(\mathfrak{b}) restricts to a section δi:[si,si+1]Xii+1\delta_{i}:[s_{i},s_{i+1}]\rightarrow X_{i}^{i+1} of f|Xii+1[si,si+1]f|_{X_{i}^{i+1}}\rightarrow[s_{i},s_{i+1}] for every i=0,,n1i=0,\ldots,n-1. Let

ΓCi:π0(Γf|Xii+1(𝔟|[si,si+1]))π1(X+(i),𝔟(s+(i)))\Gamma C_{i}:\pi_{0}(\Gamma f|_{X_{i}^{i+1}}(\mathfrak{b}|_{[s_{i},s_{i+1}]}))\rightarrow\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)}))

be the section collapse map defined in (6). By Lemma 3.8, each ΓCi\Gamma C_{i} is a bijection. This implies the following characterization of π0(Γf(𝔟))\pi_{0}(\Gamma f(\mathfrak{b})) in terms of the fundamental groups of the regular fibers of ff.

Proposition 3.12.

The product of section collapse maps

ΓC:π0(Γf(𝔟))\displaystyle\Gamma C:\pi_{0}(\Gamma f(\mathfrak{b})) i=0n1π1(X+(i),𝔟(s+(i)))\displaystyle\rightarrow\prod_{i=0}^{n-1}\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)}))
([δ0],,[δn1])\displaystyle([\delta_{0}],\ldots,[\delta_{n-1}]) (ΓC0([δ0]),,ΓCn1([δn1]))\displaystyle\mapsto(\Gamma C_{0}([\delta_{0}]),\ldots,\Gamma C_{n-1}([\delta_{n-1}]))

is bijective.

Proof.

Each map ΓCi\Gamma C_{i} for i=0,,n1i=0,\ldots,n-1 is bijective by Lemma 3.8. ∎

There is a group action

^:i=0nπ1(Xi,𝔟(si))×i=0n1π1(X+(i),𝔟(s+(i)))i=0n1π1(X+(i),𝔟(s+(i)))\hat{\star}:\prod_{i=0}^{n}\pi_{1}(X_{i},\mathfrak{b}(s_{i}))\times\prod_{i=0}^{n-1}\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)}))\rightarrow\prod_{i=0}^{n-1}\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)}))

defined as follows. The collapse maps CiC_{i} from (5) induce maps

π1(Xi,𝔟(si))\displaystyle\pi_{1}(X_{i},\mathfrak{b}(s_{i})) π1(X+(i),𝔟(s+(i))),\displaystyle\rightarrow\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)})),
π1(Xi+1,𝔟(si+1))\displaystyle\pi_{1}(X_{i+1},\mathfrak{b}(s_{i+1})) π1(X+(i),𝔟(s+(i))).\displaystyle\rightarrow\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)})).

The action ^\hat{\star} is defined for τ:=(τ0,,τn)i=0nπ1(Xi,𝔟(si))\tau:=(\tau_{0},\ldots,\tau_{n})\in\prod_{i=0}^{n}\pi_{1}(X_{i},\mathfrak{b}(s_{i})) and η:=(η0,,ηn1)i=0n1π1(X+(i),𝔟(s+(i)))\eta:=(\eta_{0},\ldots,\eta_{n-1})\in\prod_{i=0}^{n-1}\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)})) by

τ^η:=(C0(τ0)η0C0(τ1)1,,Cn1(τn1)ηn1Cn1(τn)1).\tau\,\,\hat{\star}\,\,\eta:=(C_{0}(\tau_{0})\cdot\eta_{0}\cdot C_{0}(\tau_{1})^{-1},\ldots,C_{n-1}(\tau_{n-1})\cdot\eta_{n-1}\cdot C_{n-1}(\tau_{n})^{-1}).

The actions \star and ^\hat{\star} are identified by the bijection ΓC\Gamma C from Proposition 3.12, as we now prove.

Proposition 3.13.

For τi=0nπ1(Xi,𝔟(si))\tau\in\prod_{i=0}^{n}\pi_{1}(X_{i},\mathfrak{b}(s_{i})) and δΓf(𝔟)\delta\in\Gamma f(\mathfrak{b}),

ΓC(τ[δ])=τ^ΓC([δ]).\Gamma C(\tau\star[\delta])=\tau\,\,\hat{\star}\,\,\Gamma C([\delta]).
Proof.

Write τ=(τ0,,τn)\tau=(\tau_{0},\ldots,\tau_{n}) and [δ]=([δ0],,[δn1])[\delta]=([\delta_{0}],\ldots,[\delta_{n-1}]). By definition of ΓC\Gamma C and \star, we have

ΓC(τ[δ])=(ΓC0(τ0#[δ0]#+τ11),,ΓCn1(τn1#[δn1]#+τn1)),\Gamma C(\tau\star[\delta])=\big{(}\Gamma C_{0}(\tau_{0}\#_{-}[\delta_{0}]\#_{+}\tau_{1}^{-1}),\ldots,\Gamma C_{n-1}(\tau_{n-1}\#_{-}[\delta_{n-1}]\#_{+}\tau_{n}^{-1})\big{)},

which by Lemma 3.7 is equal to

(C0(τ0)ΓC0([δ0])C0(τ1)1,,Cn1(τn1)ΓCn1([δn1])Cn1(τn)1).\big{(}C_{0}(\tau_{0})\cdot\Gamma C_{0}([\delta_{0}])\cdot C_{0}(\tau_{1})^{-1},\ldots,C_{n-1}(\tau_{n-1})\cdot\Gamma C_{n-1}([\delta_{n-1}])\cdot C_{n-1}(\tau_{n})^{-1}\big{)}.

The above expression is equal to τ^ΓC([δ])\tau\,\,\hat{\star}\,\,\Gamma C([\delta]) by definition of ^\hat{\star} and ΓC\Gamma C. ∎

The result of the above discussion is the following theorem.

Theorem 3.14.

Let f:XI=[0,1]f:X\rightarrow I=[0,1] be a tame function, ZXZX the associated diagram (4) of regular fibers, and Γf\Gamma f the space of sections. Then the natural map Π0:π0(Γf)limπ0(ZX)\Pi_{0}:\pi_{0}(\Gamma f)\rightarrow\varprojlim\pi_{0}(ZX) is surjective.

The fibers of Π0\Pi_{0} are characterized as follows. Let Ψlimπ0(ZX)\Psi\in\varprojlim\pi_{0}(ZX) and 𝔟Γf\mathfrak{b}\in\Gamma f such that Π0(𝔟)=Ψ\Pi_{0}(\mathfrak{b})=\Psi. Then Π01(Ψ)\Pi_{0}^{-1}(\Psi) is naturally in bijection with the orbits of the action ^\hat{\star} of the group i=0nπ1(Xi,𝔟(si))\prod_{i=0}^{n}\pi_{1}(X_{i},\mathfrak{b}(s_{i})) on the set i=0n1π1(X+(i),𝔟(s+(i)))\prod_{i=0}^{n-1}\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)})). \square

We now state a similar theorem when the domain of ff is S1S^{1} instead of I=[0,1]I=[0,1]. The proof is essentially the same.

Theorem 3.15.

Let XX be a compact manifold with boundary and f:XS1f:X\rightarrow S^{1} a smooth function that is submersive and such that its restriction to the boundary f|X:XS1f|_{\partial X}:\partial X\rightarrow S^{1} has isolated critical points with distinct critical values. Define the diagram ZXZX as in (4) with the additional identity map X0=XnX_{0}=X_{n}. Define the action ^\hat{\star} as above for group elements satisfying τ0=τn\tau_{0}=\tau_{n}. Then the statements in Theorem 3.14 hold. \square

4. Application: Evasion paths in mobile sensor networks

Given a collection of continuous sensors 𝒮={γ:[0,1]𝒟}{\mathcal{S}}=\{\gamma:[0,1]\rightarrow{\mathcal{D}}\} moving in a bounded domain 𝒟d{\mathcal{D}}\subset\mathbb{R}^{d}, an evasion path is a continuous intruder δ:[0,1]𝒟\delta:[0,1]\rightarrow{\mathcal{D}} that avoids detection by the sensors for the whole time interval I=[0,1]I=[0,1]. Suppose each sensor γ\gamma observes a ball of fixed radius in d\mathbb{R}^{d} centered at γ(t)\gamma(t) at every tIt\in I, and let CtC_{t} be the time-varying union of these sensor balls. Then the intruder δ\delta avoids detection if δ(t)\delta(t) is not in CtC_{t} for all tIt\in I.

The sensor ball evasion path problem (see §4.1 for details) asks for a criterion that determines whether or not an evasion path exists and that is based on the least amount of sensed information possible; for example, a common assumption is that the sensors only detect other nearby sensors and intruders; in particular, they do not know their coordinates in d\mathbb{R}^{d}. We ask to understand the homotopy type of the space {\mathcal{E}} of evasion paths. The existence problem simply asks if {\mathcal{E}} is empty or not.

We consider an idealized version of the evasion path problem, called the smoothed evasion path problem, where we smooth the time-varying covered region CtC_{t} into a smooth cobordism of manifolds with boundary; see §4.2 for the precise description. In our main result (Theorem 4.5), we establish a necessary and sufficient condition for existence of an evasion path based on the time-varying d1d-1 homology of the covered region CtC_{t} and the time-varying cup-product on cohomology H0H^{0} of its boundary, and moreover we provide a lower bound on the number of connected components π0()\pi_{0}({\mathcal{E}}) of the space of evasion paths {\mathcal{E}}. In the preliminary Corollary 4.2, we provide a full computation of π0()\pi_{0}({\mathcal{E}}) in terms of the time-varying π0\pi_{0} and π1\pi_{1} of the uncovered region Xt=𝒟CtX_{t}={\mathcal{D}}\setminus C_{t}, following from Theorem 3.14. In the case that CtC_{t} is connected for all tIt\in I, Theorem 4.5 has the simpler form Corollary 1.5 that requires only the cup product on H0H^{0} of the boundary of CtC_{t}.

Our results are the first to compute more about the space {\mathcal{E}} than whether or not it is empty.

We recall the sensor ball evasion path problem in detail and describe prior work on evasion problems in §4.1. Then we introduce the smoothed evasion problem in §4.2, and explain our results in §4.3.

4.1. Prior work

We describe the sensor ball evasion path problem in more detail. Then we explain prior results on this problem and a general evasion problem studied in [7].

Fix some sensing radius r>0r>0 and say that each sensor γ𝒮\gamma\in{\mathcal{S}} can detect objects within the closed ball

Bγ(t)={xd||γ(t)x|r}B_{\gamma(t)}=\{x\in\mathbb{R}^{d}\,\,|\,\,|\gamma(t)-x|\leq r\}

at all times tI=[0,1]t\in I=[0,1]. The time-tt covered region

(9) Ct=γ𝒮Bγ(t)dC_{t}=\bigcup_{\gamma\in{\mathcal{S}}}B_{\gamma(t)}\subset\mathbb{R}^{d}

is the union of the sensor balls, and it is homotopy equivalent to the Čech complex of the covering of CtC_{t} by the sensor balls. The time-tt uncovered region is the complement

Xt=𝒟Ct,X_{t}={\mathcal{D}}\setminus C_{t},

where 𝒟d{\mathcal{D}}\subset\mathbb{R}^{d} is a bounded domain that is homeomorphic to a ball.

Assume that the boundary 𝒟\partial{\mathcal{D}} is covered by a collection of immobile fence sensors F𝒮𝒮F{\mathcal{S}}\subset{\mathcal{S}}, meaning γ(t)d\gamma(t)\in\mathbb{R}^{d} is constant for γF𝒮\gamma\in F{\mathcal{S}}, and that the union of balls Bγ(t)B_{\gamma(t)} for γF𝒮\gamma\in F{\mathcal{S}} covers the boundary 𝒟\partial{\mathcal{D}} and is homotopy equivalent to 𝒟\partial{\mathcal{D}}. In particular, this ensures that an intruder δ\delta can never escape from the domain 𝒟{\mathcal{D}}.

We define the covered region

(10) C:=tICt×{t}d×IC:=\bigcup_{t\in I}C_{t}\times\{t\}\subset\mathbb{R}^{d}\times I

and the uncovered region

X:=(𝒟×I)C=tIXt×{t}.X:=({\mathcal{D}}\times I)\setminus C=\bigcup_{t\in I}X_{t}\times\{t\}.

Then an evasion path is equivalent to a continuous section δ:IX\delta:I\rightarrow X of the projection ρX:XI\rho_{X}:X\rightarrow I, i.e., ρXδ=idI\rho_{X}\circ\delta=id_{I}.

The sensor ball evasion path problem asks for a criterion for existence of an evasion path.

This problem was first stated and studied in [6] by de Silva and Ghrist in dimension d=2d=2. Provided that the time-varying Čech complex changes only at finitely many times, the authors establish a necessary condition for existence of an evasion path which states that the connecting homomorphism on a relative homology group with respect to the fence of the covered region must vanish.

In [1], Adams and Carlsson consider the same evasion problem in arbitrary dimension d2d\geq 2, and they establish a necessary condition for the existence of an evasion path based on zigzag persistent homology [5] of the time-varying Čech complex. They also explain how this result is equivalent to a generalization of de Silva and Ghrist’s result to the general case d2d\geq 2.

In both de Silva-Ghrist and Adams-Carlsson, the necessary condition for existence of an evasion path is determined and easily computable from the overlap information of sensor balls. Essentially, sensors only need to know which other sensors are nearby; they don’t need to know their coordinates in d\mathbb{R}^{d}.

Adams-Carlsson also study smarter sensors in the plane 2\mathbb{R}^{2} that know local distances to nearby sensors and the natural counterclockwise ordering on nearby sensors. For these smarter sensors in 2\mathbb{R}^{2}, they derive a necessary and sufficient condition for existence of an evasion path as well as an algorithm to compute it, under the assumption that the covered region CtC_{t} is connected for all tIt\in I. They show that this connectedness assumption is necessary for their methods.

In [7], Ghrist and Krishnan define positive homology and cohomology for directed spaces over q\mathbb{R}^{q}, and they compute it using sheaf-theoretic techniques. They consider a general type of evasion problem where the covered region is a pro-object in a category of smooth compact cobordisms. They derive a criterion on positive cohomology that is necessary and sufficient for existence of an evasion path, under the assumption that CtC_{t} is connected.

4.2. The smoothed evasion path problem

In this section we describe the precise version of the evasion path problem considered in this paper and indicate how our setting arises as a limiting case of the sensor ball evasion path problem as the number and density of the sensors becomes large.

Fix a dimension333We assume d2d\geq 2 because the d=0,1d=0,1 cases are trivial under the conditions of our Theorem 4.5 and would require slightly modified arguments to state as part of the theorem. See Remark 4.10 for the d=0,1d=0,1 cases. d2d\geq 2 and let 𝒟d{\mathcal{D}}\subset\mathbb{R}^{d} be a smoothly embedded closed dd-dimensional ball. The covered region C𝒟×IC\subset{\mathcal{D}}\times I, where I=[0,1]I=[0,1], is any subset with the following properties:

  • The time 0 and 11 covered regions, given by

    Ci=C(d×{i})C_{i}=C\cap(\mathbb{R}^{d}\times\{i\})

    for i=0,1i=0,1, are smooth compact codimension-0 submanifolds CidC_{i}\subset\mathbb{R}^{d} with smooth boundary Ci\partial C_{i}.

  • The full covered region Cd×IC\subset\mathbb{R}^{d}\times I is a smooth embedded compact cobordism (Definition 2.1) between C0C_{0} and C1.C_{1}. In particular, CC has boundary

    C=C0C1\partial C=C_{0}\cup\partial\cup C_{1}

    where \partial is a dd-dimensional manifold with boundary C0C1\partial C_{0}\sqcup\partial C_{1} that intersects CiC_{i} along its boundary

    Ci=Ci\partial\cap C_{i}=\partial C_{i}

    for i=0,1i=0,1.

  • The boundary \partial contains 𝒟×I\partial{\mathcal{D}}\times I as a connected component.

  • The critical points of the projection ρ:I\rho_{\partial}:\partial\rightarrow I are isolated and have distinct critical values contained in (0,1)(0,1).

Remark 4.1.

Heuristically, when the covered region CC is given by the union of sensor balls over all times tIt\in I as in (9) and (10), the smoothed evasion path problem arises from a small perturbation of CC inside d×I\mathbb{R}^{d}\times I that smooths out the time-varying union of sensor balls into a manifold. The projection CIC\rightarrow I does not have any critical points since CC is codimension-0 in d×I\mathbb{R}^{d}\times I, and generically the projection on the boundary ρ:I\rho_{\partial}:\partial\rightarrow I has isolated critical points with distinct critical values.

For generic sensor paths in d\mathbb{R}^{d}, the boundaries of the sensors balls will intersect transversely at all but finitely many times tIt\in I. We call the transverse times the regular times, and the non-transverse times are the critical times. Heuristically, these times correspond to the regular values and the critical values, respectfully, of the function ρ\rho_{\partial} in the smoothed evasion path problem.

We define the uncovered region XX to be the closure of the complement (𝒟×I)C({\mathcal{D}}\times I)\setminus C. More precisely, we define the essential boundary

B:=(𝒟×I),B:=\partial\setminus(\partial{\mathcal{D}}\times I),

and the uncovered region

X:=((𝒟×I)C)B.X:=(({\mathcal{D}}\times I)\setminus C)\cup B.

Then XX is a compact cobordism (Definition 2.1) between the time 0 and time 11 uncovered regions Xi=X(d×{i})X_{i}=X\cap(\mathbb{R}^{d}\times\{i\}) for i=0,1,i=0,1, with boundary

X=X0BX1.\partial X=X_{0}\cup B\cup X_{1}.

Note that BB is the intersection

B=XC.B=X\cap C.

Consider the projection ρ:d×II\rho:\mathbb{R}^{d}\times I\rightarrow I and the space of evasion paths

(11) :={δ:IX|ρδ=idI and δ is continuous},{\mathcal{E}}:=\{\delta:I\rightarrow X\,\,|\,\,\rho\circ\delta=id_{I}\text{ and }\delta\text{ is continuous}\},

i.e., the space of continuous sections of the projection restricted to the uncovered region

ρX:XI.\rho_{X}:X\rightarrow I.

Here, {\mathcal{E}} is given the compact-open topology, or equivalently the topology of the supremum norm for continuous maps IdI\rightarrow\mathbb{R}^{d} with respect to the standard norm on d\mathbb{R}^{d}.

Our main result (Theorem 4.5) establishes a necessary and sufficient condition for existence of an evasion path, and moreover provides a lower bound on the number of connected components π0()\pi_{0}({\mathcal{E}}). The required information consists of a parameterized homology of the covered region CC, a parameterized cup-product on the essential boundary BB, and an Alexander duality isomorphism. See also Corollary 1.5 for the simpler case when CtC_{t} is connected for all tt.

We now briefly discuss how one might approximate the information required in Theorem 4.5 to compute existence of an evasion path in the sensor ball evasion path problem where the covered region CC is a time-varying union of sensors balls as in (9) and (10).

To compute the time-varying homology of CC it suffices to have the information of the time-varying Čech complex, or in other words the information of sensor ball overlaps; see [1].

Heuristically, one can use the Niyogi-Smale-Weinberger Theorem [13] to compute the time-varying cup-product on cohomology H0H^{0} of the essential boundary BB under the following additional assumptions:

  • Directional sensing: The sensors detect the direction in d\mathbb{R}^{d} of nearby sensors, as well as the direction of the wall 𝒟\partial{\mathcal{D}} if it is nearby.

  • Dense coverage: The sensors are ϵ\epsilon-densely distributed throughout CtC_{t} at all regular values tt, in the sense of [13].

Indeed, with these assumptions, we can label a sensor γ𝒮\gamma\in{\mathcal{S}} at time tIt\in I as an essential boundary sensor if there is a codimension-11 hypersurface in d\mathbb{R}^{d} containing γ(t)\gamma(t) such that all sensors in a small neighborhood of γ(t)\gamma(t) lie on the same side of the hypersurface, and such that the other side of the hypersurface is not the wall 𝒟\partial{\mathcal{D}}. First of all, sensors can detect this information due to the directional sensing hypothesis. Second, due to the dense coverage hypothesis, such a hypersurface exists if and only if the sensor is close to the essential boundary BB; indeed, sensors in the interior of CC see other sensors in all directions, whereas those near BB see other sensors towards the interior of CC and they don’t see any sensors on the other side of BB. This same effect occurs near 𝒟\partial{\mathcal{D}}, yet the sensor is aware that it is an effect of the wall. The Niyogi-Smale-Weinberger Theorem [13] then asserts that the Čech complex of the boundary sensors computes cohomology of Bt=B(d×{t})B_{t}=B\cap(\mathbb{R}^{d}\times\{t\}) at regular times tt.

4.3. Connected components of the space of evasion paths

In this section we prove the results, Theorem 4.5, Corollary 4.2, and Corollary 1.5, about the smoothed evasion path problem introduced in §4.2.

Denote the critical values of the projection ρB:BI\rho_{B}:B\rightarrow I by

t1<t2<<tnt_{1}<t_{2}<\cdots<t_{n}

and choose interleaving values

0=s0<t1<s1<t2<<sn1<tn<sn=1.0=s_{0}<t_{1}<s_{1}<t_{2}<\cdots<s_{n-1}<t_{n}<s_{n}=1.

We recall the zigzag discretization procedure described in §2.2, which we apply here to various subspaces Md×I.M\subset\mathbb{R}^{d}\times I. Consider the projection

ρ:d×II\rho:\mathbb{R}^{d}\times I\rightarrow I

and its restriction

ρM:=ρ|M:MI,\rho_{M}:=\rho|_{M}:M\rightarrow I,

from which we obtain subspaces of MM given by the preimages

Mi\displaystyle M_{i} :=ρM1(si), for 0in,\displaystyle:=\rho_{M}^{-1}(s_{i}),\text{ for }0\leq i\leq n,
Mii+1\displaystyle M_{i}^{i+1} :=ρM1([si,si+1]), for 0in1.\displaystyle:=\rho_{M}^{-1}([s_{i},s_{i+1}]),\text{ for }0\leq i\leq n-1.

These fit together into a zigzag diagram of topological spaces

ZM~:=(M0M01M1M12Mn1nMn)\widetilde{ZM}:=\bigg{(}M_{0}\hookrightarrow M_{0}^{1}\hookleftarrow M_{1}\hookrightarrow M_{1}^{2}\hookleftarrow\cdots\hookrightarrow M_{n-1}^{n}\hookleftarrow M_{n}\bigg{)}

where all maps are the natural inclusions.

Consider now the zigzag diagram ZX~\widetilde{ZX} in the case that M=XM=X is the uncovered region. Recall the abbreviated zigzag diagram

ZX:=(X0X1Xn)ZX:=(X_{0}\leftrightarrow X_{1}\leftrightarrow\cdots\leftrightarrow X_{n})

defined in (4), which is obtained roughly by inverting homotopy equivalences. Note that

limπ0(ZX~)=limπ0(ZX).\varprojlim\pi_{0}(\widetilde{ZX})=\varprojlim\pi_{0}(ZX).

Theorem 3.14 applied to the projection ρX:XI\rho_{X}:X\rightarrow I provides the following.

Corollary 4.2.

There is a surjection

Π0:π0()limπ0(ZX)=limπ0(ZX~).\Pi_{0}:\pi_{0}({\mathcal{E}})\rightarrow\varprojlim\pi_{0}(ZX)=\varprojlim\pi_{0}(\widetilde{ZX}).

In the notation of §3, the fibers of Π0\Pi_{0} are characterized as follows. Let Ψlimπ0(ZX)\Psi\in\varprojlim\pi_{0}(ZX) and 𝔟\mathfrak{b}\in{\mathcal{E}} such that Π0(𝔟)=Ψ\Pi_{0}(\mathfrak{b})=\Psi. Then the fiber Π01(Ψ)\Pi_{0}^{-1}(\Psi) is naturally in bijection with the orbits of the action ^\hat{\star} of the group i=0nπ1(Xi,𝔟(si))\prod_{i=0}^{n}\pi_{1}(X_{i},\mathfrak{b}(s_{i})) on the set i=0n1π1(X+(i),𝔟(s+(i)))\prod_{i=0}^{n-1}\pi_{1}(X_{\partial^{+}(i)},\mathfrak{b}(s_{\partial^{+}(i)})).

Proof.

The space {\mathcal{E}}, defined in (11), is the space of continuous sections of the projection ρX:XI\rho_{X}:X\rightarrow I. Since Xd×IX\subset\mathbb{R}^{d}\times I is codimension-0, the projection ρX\rho_{X} is submersive. Since the projection on the boundary ρ:I\rho_{\partial}:\partial\rightarrow I has isolated critical points with distinct critical values in (0,1)(0,1) by hypothesis, the same is true for its restriction to the essential boundary BB\subset\partial, and hence ρX\rho_{X} is a tame function (Definition 2.2). Hence the result follows from Theorem 3.14. ∎

Remark 4.3.

Corollary 4.2, applied to the examples presented in Figure 1 produces the same full computation of π0()\pi_{0}({\mathcal{E}}) as Theorem 1.1, namely exactly that described in Remark 1.2.

Our main result, Theorem 4.5, provides only lower bounds on the cardinality of π0()\pi_{0}({\mathcal{E}}) and determines if it is empty or not, however it needs as input only (co)homological information about the covered region CC and the essential boundary BB. This is important for applications in which the topology of CC and the boundary can be detected from the sensor information but the topology of XX may be harder to determine since this is defined as the region where there are no sensors. To improve this to a full computation of π0()\pi_{0}({\mathcal{E}}), one would need to compute the fiberwise fundamental groups and maps between them used in Corollary 4.2 from homological information. This can be addressed by a fiberwise version of the unstable Adams spectral sequence of Bousfield-Kan, currently under development by Wyatt Mackey.

The goal now is to compute π0(ZX~)\pi_{0}(\widetilde{ZX}) in terms of homological information about the covered region CC and the boundary BB. In principle, we have access to this homological information in applications with moving sensors (see the end of §4.2).

Let kk be a field. In view of Proposition 4.4 below, to compute π0(ZX~)\pi_{0}(\widetilde{ZX}) it suffices to compute the cup product structure on H0(ZX~;k)H^{0}(\widetilde{ZX};k).

Let Homkalgebra(,k)Hom_{k-algebra}(-,k) be the functor that takes a kk-algebra RR to the set of kk-algebra homomorphisms RkR\rightarrow k. Then applying Homkalgebra(,k)Hom_{k-algebra}(-,k) to the diagram H0(ZX~;k)H^{0}(\widetilde{ZX};k) yields a zigzag diagram of sets.

Proposition 4.4.

There is an isomorphism of zigzag diagrams of sets

π0(ZX~)Homkalgebra(H0(ZX~;k),k),\pi_{0}(\widetilde{ZX})\cong Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k),

or in other words a commutative diagram

π0(X0){\pi_{0}({X}_{0})}π0(X01){\pi_{0}({X}_{0}^{1})}{\cdots}π(Xn1n){\pi({X}_{n-1}^{n})}π0(Xn){\pi_{0}({X}_{n})}H0(X0;k){H^{0}({X}_{0};k)^{\vee}}H0(X01;k){H^{0}({X}_{0}^{1};k)^{\vee}}{\cdots}H0(Xn1n;k){H^{0}({X}_{n-1}^{n};k)^{\vee}}H0(Xn;k){H^{0}({X}_{n};k)^{\vee}}

with all vertical arrows bijections and where H0(,k):=Homkalgebra(H0(;k),k)H^{0}(-,k)^{\vee}:=Hom_{k-algebra}(H^{0}(-;k),k) is the set of kk-algebra morphisms H0(;k)kH^{0}(-;k)\rightarrow k with respect to the cup product on H0H^{0}.

Proof.

For any topological space AA, there is a bijection of sets

α:π0(A)\displaystyle\alpha:\pi_{0}(A) Homkalgebra(H0(A;k),k)\displaystyle\rightarrow Hom_{k-algebra}(H^{0}(A;k),k)
[a]\displaystyle[a] ((φH0(A;k))φ(a)k).\displaystyle\mapsto\bigg{(}(\varphi\in H^{0}(A;k))\mapsto\varphi(a)\in k\bigg{)}.

Moreover, α\alpha is a natural isomorphism from the functor π0()\pi_{0}(-) to the functor Homkalgebra(H0(;k),k)Hom_{k-algebra}(H^{0}(-;k),k). Both of these are functors from the category of topological spaces to the category of sets. Hence α\alpha induces the claimed isomorphism of zigzag diagrams. ∎

It remains to compute H0(ZX~;k)H^{0}(\widetilde{ZX};k). The inclusion map ιB:BX\iota_{B}:B\rightarrow X induces a map of zigzag diagrams of kk-algebras ιB:H0(ZX~;k)H0(ZB~;k)\iota_{B}^{*}:H^{0}(\widetilde{ZX};k)\rightarrow H^{0}(\widetilde{ZB};k). In the simplest situation, when the time-tt covered region C(d×{t})C\cap(\mathbb{R}^{d}\times\{t\}) is connected for all tIt\in I, the map ιB\iota_{B}^{*} is an isomorphism by Proposition 4.7 below. The final result in this case is Corollary 1.5.

In Theorem 4.5 below, we explain how to compute H0(ZX~;k)H^{0}(\widetilde{ZX};k) in the case that C(d×{t})C\cap(\mathbb{R}^{d}\times\{t\}) is not necessarily connected. Let

Bc=(𝒟×I)BB^{c}=({\mathcal{D}}\times I)\setminus B

be the complement of the essential boundary. Recall the notation for the regular level sets Bi=B(d×{si}),Bic=Bc(d×{si})B_{i}=B\cap(\mathbb{R}^{d}\times\{s_{i}\}),B^{c}_{i}=B^{c}\cap(\mathbb{R}^{d}\times\{s_{i}\}), and Ci=C(d×{si})C_{i}=C\cap(\mathbb{R}^{d}\times\{s_{i}\}) for 0in0\leq i\leq n. For each 0in0\leq i\leq n, there is a composite map

Hd1(Ci;k)ιCiHd1(Bic;k)αiH0(Bi;k)H_{d-1}(C_{i};k)\xrightarrow{{\iota_{C_{i}}}_{*}}H_{d-1}(B^{c}_{i};k)\xrightarrow{\alpha_{i}}H^{0}(B_{i};k)

where ιCi{\iota_{C_{i}}}_{*} is induced by the inclusion444For a manifold CC with boundary, the inclusion Interior(C)C\text{Interior}(C)\subset C is a homotopy equivalence. ιCi:Interior(Ci)Bic\iota_{C_{i}}:\text{Interior}(C_{i})\rightarrow B_{i}^{c} and αi\alpha_{i} is Alexander duality (see Remark 4.8).

See Example 4.9 below for the computation of H0(ZX~;k)H^{0}(\widetilde{ZX};k) in example (a) from Figure 1.

Theorem 4.5.

There is a surjection π0()limHomkalgebra(H0(ZX~;k),k)\pi_{0}({\mathcal{E}})\rightarrow\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k). In particular, an evasion path exists (i.e. {\mathcal{E}} is nonempty) if and only if limHomkalgebra(H0(ZX~;k),k)\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k) is nonempty, and the cardinality of π0()\pi_{0}({\mathcal{E}}) is bounded from below by the cardinality of the inverse limit.

Assume that the projection CIC\rightarrow I does not have any local maxima or local minima except over 0,1I0,1\in I. Then the following information determines the zigzag diagram of kk-algebras H0(ZX~;k)H^{0}(\widetilde{ZX};k) up to isomorphism:

  1. (i)

    The zigzag diagram of kk-algebras H0(ZB~;k)H^{0}(\widetilde{ZB};k),

  2. (ii)

    The images im(αiιCi)H0(Bi;k)im(\alpha_{i}\circ{\iota_{C_{i}}}_{*})\subset H^{0}(B_{i};k) for 0in0\leq i\leq n,

  3. (iii)

    The type (N or D) of every boundary critical point of CIC\rightarrow I.

Remark 4.6.

The assumption and the given information (i)(iii)(i)-(iii) in Theorem 4.5 are justified when interpreted in the context of a mobile sensor network.

Indeed, a local maximum of CIC\rightarrow I can only occur if a sensor disappears. A local minimum can occur only if a new sensor is created. So, the assumption that CIC\rightarrow I does not have local minima or maxima at intermediate times t(0,1)t\in(0,1) means that the result applies to mobile sensor networks in which sensors do not get added or removed during the time interval of interest.

The assumed information in (i)-(iii) is all determined by the boundary BB and the covered region CC. Hence, it is reasonable to assume that the sensors, which cover CC and have boundary BB, can record this information; see the discussion of the Niyogi-Smale-Weinberger theorem at the end of §4.2 for more detail. In particular, we do not require a priori information about the uncovered region XX; indeed, the theorem computes the required cohomological information H0(ZX~;k)H^{0}(\widetilde{ZX};k) from the given information in (i)(iii)(i)-(iii).

Proof of Theorem 4.5.

The claimed surjection

π0()limHomkalgebra(H0(ZX~;k),k)\pi_{0}({\mathcal{E}})\rightarrow\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k)

is the composition of the surjection Π0:π0()limπ0(ZX~)\Pi_{0}:\pi_{0}({\mathcal{E}})\rightarrow\varprojlim\pi_{0}(\widetilde{ZX}) from Corollary 4.2 and the isomorphism limπ0(ZX~)limHomkalgebra(H0(ZX~;k),k)\varprojlim\pi_{0}(\widetilde{ZX})\cong\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k) from Proposition 4.4. It remains to compute H0(ZX~;k)H^{0}(\widetilde{ZX};k) from (i)-(iii) under the assumption that CIC\rightarrow I does not have any local maxima or minima at intermediate times t(0,1)t\in(0,1).

The inclusion

ιB:BX\iota_{B}:B\rightarrow X

induces a map of zigzag diagrams of kk-algebras

ιB:H0(ZX~;k)H0(ZB~;k).\iota_{B}^{*}:H^{0}(\widetilde{ZX};k)\rightarrow H^{0}(\widetilde{ZB};k).
Proposition 4.7.

The map ιB\iota_{B}^{*} is injective. If the time-tt covered region C(d×{t})C\cap(\mathbb{R}^{d}\times\{t\}) is connected for all tIt\in I, then ιB\iota_{B}^{*} is an isomorphism.

Proof.

The claim that ιB\iota_{B}^{*} is injective means that the restriction maps

ιi:H0(Xi)\displaystyle\iota_{i}^{*}:H^{0}(X_{i}) H0(Bi),\displaystyle\rightarrow H^{0}(B_{i}),
ιii+1:H0(Xii+1)\displaystyle{\iota_{i}^{i+1}}^{*}:H^{0}(X_{i}^{i+1}) H0(Bii+1)\displaystyle\rightarrow H^{0}(B_{i}^{i+1})

induced by the inclusions BiXiB_{i}\subset X_{i} and Bii+1Xii+1B_{i}^{i+1}\subset X_{i}^{i+1} are injective for all ii.

We first consider the restriction maps ιi\iota_{i}^{*}. By definition, we have the covering XiCi=𝒟X_{i}\cup C_{i}={\mathcal{D}} and the intersection along their common manifold boundary is XiCi=BiX_{i}\cap C_{i}=B_{i}. Hence there is a Mayer-Vietoris sequence

H~0(𝒟)H~0(Xi)H~0(Ci)H~0(Bi)H~1(𝒟).\tilde{H}^{0}({\mathcal{D}})\rightarrow\tilde{H}^{0}(X_{i})\oplus\tilde{H}^{0}(C_{i})\rightarrow\tilde{H}^{0}(B_{i})\rightarrow\tilde{H}^{1}({\mathcal{D}}).

Since 𝒟{\mathcal{D}} is homeomorphic to a ball and hence contractible, it follows that the map H~0(Xi)H~0(Ci)H~0(Bi)\tilde{H}^{0}(X_{i})\oplus\tilde{H}^{0}(C_{i})\rightarrow\tilde{H}^{0}(B_{i}) is an isomorphism. Restricted to H~0(Xi){0}\tilde{H}^{0}(X_{i})\oplus\{0\}, this is exactly the restriction map on reduced cohomology induced by BiXiB_{i}\subset X_{i}, and it is injective. Hence the induced map on unreduced cohomology, which is ιi\iota_{i}^{*}, is also injective, as claimed.

Under the additional hypothesis that the time-tt covered region C(d×{t})C\cap(\mathbb{R}^{d}\times\{t\}) is connected for all tIt\in I, we have that Ci=C(d×{si})C_{i}=C\cap(\mathbb{R}^{d}\times\{s_{i}\}) is connected and so H~0(Ci)=0\tilde{H}^{0}(C_{i})=0, in which case ιi\iota_{i}^{*} is an isomorphism, as claimed.

We now return to the general case and analyze the restriction maps ιii+1{\iota_{i}^{i+1}}^{*}. The argument that these are injective is similar. By definition, we have the covering Xii+1Cii+1=𝒟×[si,si+1]X_{i}^{i+1}\cup C_{i}^{i+1}={\mathcal{D}}\times[s_{i},s_{i+1}] with intersection along the common boundary Xii+1Cii+1=Bii+1X_{i}^{i+1}\cap C_{i}^{i+1}=B_{i}^{i+1}. Hence there is a Mayer-Vietoris sequence

H~0(𝒟×[si,si+1])H~0(Xii+1)H~0(Cii+1)H~0(Bii+1)H~1(𝒟×[si×si+1]).\tilde{H}^{0}({\mathcal{D}}\times[s_{i},s_{i+1}])\rightarrow\tilde{H}^{0}(X_{i}^{i+1})\oplus\tilde{H}^{0}(C_{i}^{i+1})\rightarrow\tilde{H}^{0}(B_{i}^{i+1})\rightarrow\tilde{H}^{1}({\mathcal{D}}\times[s_{i}\times s_{i+1}]).

Hence the map H~0(Xii+1)H~0(Cii+1)H~0(Bii+1)\tilde{H}^{0}(X_{i}^{i+1})\oplus\tilde{H}^{0}(C_{i}^{i+1})\rightarrow\tilde{H}^{0}(B_{i}^{i+1}) is an isomorphism, which implies that the restriction map ιii+1{\iota_{i}^{i+1}}^{*} is injective, as claimed.

Under the additional hypothesis that the time-tt covered region C(d×{t})C\cap(\mathbb{R}^{d}\times\{t\}) is connected for all tIt\in I, it follows from Lemma 2.6 applied to the tame function Cii+1IC_{i}^{i+1}\rightarrow I that the cobordism Cii+1C_{i}^{i+1} is homotopy equivalent to either CiC_{i} or Ci+1C_{i+1}, hence is connected, for all ii. Thus H~0(Cii+1)=0\tilde{H}^{0}(C_{i}^{i+1})=0 and so ιii+1{\iota_{i}^{i+1}}^{*} is an isomorphism, as claimed. This completes the proof of the proposition. ∎

Proposition 4.7 implies that H0(ZX~;k)H^{0}(\widetilde{ZX};k) is isomorphic to the image im(ιB)\text{im}(\iota_{B}^{*}) of ιB\iota_{B}^{*} a zigzag diagram of kk-algebras, where the zigzag kk-algebra structure on im(ιB)\text{im}(\iota_{B}^{*}) is induced by the cup-product on H0(ZB~;k)H^{0}(\widetilde{ZB};k). Hence, it suffices to compute

im(ιBi)H0(Bi;k) and im(ιBii+1)H0(Bii+1;k)\text{im}(\iota_{B_{i}}^{*})\subset H^{0}(B_{i};k)\,\,\text{ and }\,\,\text{im}(\iota_{B_{i}^{i+1}}^{*})\subset H^{0}(B_{i}^{i+1};k)

as kk-vector spaces for all ii. Indeed, the kk-algebra structures on these images are induced by the cup products on H0(Bi;k)H^{0}(B_{i};k) and H0(Bii+1;k)H^{0}(B_{i}^{i+1};k), respectively, which are given by assumption (i)(i).

For each ii, there is a commutative diagram of kk-vector spaces

H0(Xi;k){H^{0}(X_{i};k)}H0(Bi;k){H^{0}(B_{i};k)}Hd1(Ci;k){H_{d-1}(C_{i};k)}Hd1(Bic;k),{H_{d-1}(B_{i}^{c};k),}ιBi\scriptstyle{\iota_{B_{i}}^{*}}𝒜i\scriptstyle{{\mathcal{A}}_{i}}αi1\scriptstyle{\alpha_{i}^{-1}}ιCi\scriptstyle{{\iota_{C_{i}}}_{*}}

where the top map is the restriction map on cohomology induced by the inclusion ιBi:BiXi\iota_{B_{i}}:B_{i}\hookrightarrow X_{i}, the bottom map is the pushforward on homology Hd1(;k)H_{d-1}(-;k) induced by the inclusion ιCi:Interior(Ci)Bc\iota_{C_{i}}:Interior(C_{i})\rightarrow B^{c}, and the vertical maps 𝒜i{\mathcal{A}}_{i} and αi\alpha_{i} are Alexander duality isomorphisms (see Remark 4.8). This diagram commutes, and hence

im(ιBi)=im(αiιCi).im(\iota^{*}_{B_{i}})=im(\alpha_{i}\circ{\iota_{C_{i}}}_{*}).

The right hand side is given by assumption (ii)(ii).

Remark 4.8.

In the Alexander duality isomorphisms 𝒜i{\mathcal{A}}_{i} and αi\alpha_{i} we have unreduced cohomology H0H^{0} instead of reduced cohomology H~0\tilde{H}^{0} because XiCi=𝒟X_{i}\cup C_{i}={\mathcal{D}} and BiBic=𝒟B_{i}\cup B_{i}^{c}={\mathcal{D}} are missing the connected subset d𝒟\mathbb{R}^{d}\setminus{\mathcal{D}} of d\mathbb{R}^{d} and moreover d𝒟\mathbb{R}^{d}\setminus{\mathcal{D}} is disjoint from both XiX_{i} and BiB_{i}.

It remains to compute im(ιBii+1).im(\iota^{*}_{B_{i}^{i+1}}). Consider the following commutative diagram, which is the relevant part of the map H0(ZX~)H0(ZB~)H^{0}(\widetilde{ZX})\rightarrow H^{0}(\widetilde{ZB}) induced by the inclusion BXB\hookrightarrow X,

H0(Xi){H^{0}(X_{i})}H0(Xii+1){H^{0}(X_{i}^{i+1})}H0(Xi+1){H^{0}(X_{i+1})}H0(Bi){H^{0}(B_{i})}H0(Bii+1){H^{0}(B_{i}^{i+1})}H0(Bi+1).{H^{0}(B_{i+1}).}ιBi\scriptstyle{\iota_{B_{i}}^{*}}ϕ\scriptstyle{\phi}ιBii+1\scriptstyle{\iota_{B_{i}^{i+1}}^{*}}ιBi+1\scriptstyle{\iota_{B_{i+1}}^{*}}φ\scriptstyle{\varphi}

All vertical maps are injective by Proposition 4.7.

There is a single critical point pBii+1p\in B_{i}^{i+1} of the projection BIB\rightarrow I, and by assumption (iii)(iii) we know if pp is type-NN or type-DD with respect to CIC\rightarrow I. Assume now that it is type-NN; the argument in the type-DD case is symmetric. Since pp is type-NN with respect to the projection CIC\rightarrow I, it is type-DD with respect to the projection XIX\rightarrow I. Hence Xii+1X_{i}^{i+1} deformation retracts onto XiX_{i} by Lemma 2.6, which implies that ϕ\phi is an isomorphism.

We claim that φ\varphi is injective. Indeed, any nontrivial kernel of φ\varphi implies the existence of a component AA of the cobordism Bii+1B_{i}^{i+1} whose boundary A\partial A is disjoint from BiB_{i} and hence lies entirely in Bi+1B_{i+1}. Hence, since pp is the only critical point in Bii+1B_{i}^{i+1}, it must lie on AA, since any component of Bii+1B_{i}^{i+1} without a critical point is a trivial cobordism. It follows that pp is a local minimum of BIB\rightarrow I. Since pp is type-NN with respect to CIC\rightarrow I, the normal vector to BB at pp pointing into the interior of CC is pointing positively along II and hence pp is also a local minimum of CIC\rightarrow I. This contradicts the assumption that there are no local minima of CIC\rightarrow I away from 0 and 11, proving the claim that φ\varphi is injective.

To complete the proof, recall that in the commutative diagram above all vertical maps are injective, ϕ\phi is an isomorphism, and φ\varphi is injective. It follows that im(ιBii+1)=φ1(im(ιBi))im(\iota^{*}_{B_{i}^{i+1}})=\varphi^{-1}(im(\iota^{*}_{B_{i}})). Since we computed im(ιBi)im(\iota^{*}_{B_{i}}) above and since the map φ\varphi is part of the given data in assumption (i)(i), this completes the computation of im(ιBii+1)im(\iota^{*}_{B_{i}^{i+1}}) and the proof of Theorem 4.5. ∎

Example 4.9.

We use the method in Theorem 4.5 to compute H0(ZX~;k)H^{0}(\widetilde{ZX};k) in example (a) from Figure 1, providing the lower bound |π0()||limHomkalgebra(H0(ZX~;k),k)|=2|\pi_{0}({\mathcal{E}})|\geq|\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k)|=2.

The zigzag diagram of spaces ZB~=(B0BB1)\widetilde{ZB}=(B_{0}\hookrightarrow B\hookleftarrow B_{1}) is a circle B0=S1B_{0}=S^{1} and two circles B1=S1S1B_{1}=S^{1}\sqcup S^{1} included as the boundary of the pair of pants BB. Hence we have the zigzag diagram of kk-algebras

H0(ZB~;k)=(kkkk).\displaystyle H^{0}(\widetilde{ZB};k)=(k\hookleftarrow k\hookrightarrow k\oplus k).

Since the time-tt covered region is connected for all tIt\in I, the map ιB:H0(ZX~;k)H0(ZB~;k)\iota_{B}^{*}:H^{0}(\widetilde{ZX};k)\rightarrow H^{0}(\widetilde{ZB};k) is an isomorphism by Proposition 4.7. Hence limHomkalgebra(H0(ZX~;k),k)\varprojlim Hom_{k-algebra}(H^{0}(\widetilde{ZX};k),k) has cardinality 22.

Remark 4.10.

Theorem 4.5 is stated and proved in dimensions d2d\geq 2. Here we explain the d=0,1d=0,1 cases. For d=0d=0, the only possibility is C=0×IC=\mathbb{R}^{0}\times I and the uncovered region is empty X=X=\emptyset so there are no evasion paths =.{\mathcal{E}}=\emptyset. In the d=1d=1 case, the assumption that the projection CIC\rightarrow I does not have any local minima or maxima away from 0,1I0,1\in I implies that the cardinality of π0()\pi_{0}({\mathcal{E}}) is equal to the number of connected components of the time-0 uncovered region π0(X0)\pi_{0}(X_{0}) minus the number of type D boundary critical points of XIX\rightarrow I. This is equal to |π0()|=|π0(C0)|1#(type N critical points of CI).|\pi_{0}({\mathcal{E}})|=|\pi_{0}(C_{0})|-1-\#(\text{type N critical points of }C\rightarrow I).

References

  • [1] H. Adams and G. Carlsson, Evasion paths in mobile sensor networks, The International Journal of Robotics Research 34 (2015), no. 1, 90–104.
  • [2] J. Bloom, The combinatorics of Morse theory with boundary, Proceedings of the Gökova Geometry-Topology Conference 2012, Int. Press, Somerville, MA, 2013, pp. 43–88.
  • [3] M. Borodzik, A. Némethi, and A. Ranicki, Morse theory for manifolds with boundary, Algebr. Geom. Topol. 16 (2016), no. 2, 971–1023.
  • [4] D. Braess, Morse-Theorie für berandete Mannigfaltigkeiten, Math. Ann. 208 (1974), 133–148.
  • [5] G. Carlsson and V. de Silva, Zigzag persistence, Found. Comput. Math. 10 (2010), no. 4, 367–405.
  • [6] V. de Silva and R. Ghrist, Coordinate-free coverage in sensor networks with controlled boundaries via homology, International Journal of Robotics Research 25 (2006), 1205–1222.
  • [7] R. Ghrist and S. Krishnan, Positive Alexander duality for pursuit and evasion, SIAM J. Appl. Algebra Geom. 1 (2017), no. 1, 308–327.
  • [8] B. Hajduk, Minimal mm-functions, Fund. Math. 111 (1981), no. 3, 179–200.
  • [9] A. Jankowski and R. Rubinsztein, Functions with non-degenerate critical points on manifolds with boundary, Comment. Math. Prace Mat. 16 (1972), 99–112.
  • [10] P. Kronheimer and T. Mrowka, Monopoles and three-manifolds, New Mathematical Monographs, vol. 10, Cambridge University Press, Cambridge, 2007.
  • [11] F. Laudenbach, A Morse complex on manifolds with boundary, Geom. Dedicata 153 (2011), 47–57.
  • [12] J. Milnor, Lectures on the hh-cobordism theorem, Notes by L. Siebenmann and J. Sondow, Princeton University Press, Princeton, N.J., 1965.
  • [13] P. Niyogi, S. Smale, and S. Weinberger, Finding the homology of submanifolds with high confidence from random samples, Discrete Comput. Geom. 39 (2008), no. 1-3, 419–441.