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

Moderately Discontinuous Homotopy

J. Fernández de Bobadilla Javier Fernández de Bobadilla: (1) IKERBASQUE, Basque Foundation for Science, Maria Diaz de Haro 3, 48013, Bilbao, Bizkaia, Spain (2) BCAM Basque Center for Applied Mathematics, Mazarredo 14, E48009 Bilbao, Basque Country, Spain (3) Academic Colaborator at UPV/EHU [email protected] S. Heinze Sonja Heinze: BCAM Basque Center for Applied Mathematics, Mazarredo 14, E48009 Bilbao, Basque Country, Spain [email protected]  and  M. Pe Pereira Facultad de Ciencias Matematicas and Instituto de Matemática Interdisciplinar, Universidad Complutense de Madrid, Plaza de Ciencias, 3, Ciudad Universitaria, 28040 MADRID [email protected]
Abstract.

We introduce a metric homotopy theory, which we call Moderately Discontinuous Homotopy, designed to capture Lipschitz properties of metric singular subanalytic germs. It matches with the Moderately Discontinuous Homology theory receantly developed by the authors and E. Sampaio. The kk-th MD homotopy group is a group MDHbMDH^{b}_{\bullet} for any b[1,]b\in[1,\infty] together with homomorphisms MDπbMDπbMD\pi^{b}\to MD\pi^{b^{\prime}} for any bbb\geq b^{\prime}. We develop all its basic properties including finite presentation of the groups, long homology sequences of pairs, metric homotopy invariance, Seifert- van Kampen Theorem and the Hurewicz isomorphism Theorem. We prove comparison theorems that allow to relate the metric homotopy groups with topological homotopy groups of associated spaces. For b=1b=1 it recovers the homotopy groups of the tangent cone for the outer metric and of the Gromov tangent cone for the inner one. In general, for b=b=\infty the MDMD- homotopy recovers the homotopy of the punctured germ. Hence, our invariant can be seen as an algebraic invariant interpolating the homotopy from the germ to its tangent cone. We end the paper with a full computation of our invariant for any normal surface singularity for the inner metric. We also provide a full computation of the MD-Homology in the same case.

2010 Mathematics Subject Classification:
Primary 14B05,32S05,32S50,55N35,51F99
The first author is supported by ERCEA 615655 NMST Consolidator Grant, MINECO by the project reference MTM2016-76868-C2-1-P (UCM), by the Basque Government through the BERC 2018-2021 program and Gobierno Vasco Grant IT1094-16, by the Spanish Ministry of Science, Innovation and Universities: BCAM Severo Ochoa accreditation SEV-2017-0718. The second author is supported by a La Caixa Ph.D. grant associated to the MINECO project SEV-2011-0087, by ERCEA 615655 NMST Consolidator Grant, MINECO by the project reference MTM2016-76868-C2-1-P (UCM), by the Basque Government through the BERC 2018-2021 program and Gobierno Vasco Grant IT1094-16, by the Spanish Ministry of Science, Innovation and Universities: BCAM Severo Ochoa accreditation SEV-2017-0718. The third author is supported by MINECO by the project reference MTM 2017-89420 and MTM2016-76868-C2-1-P (UCM).

1. Introduction

We introduce a metric homotopy theory, which we call Moderately Discontinuous Homotopy, designed to capture Lipschitz properties of metric singular subanalytic germs. This theory matches with the Moderately Discontinuous Homology theory receantly developed by the authors and E. Sampaio in [8]. Both theories run parallel to the classical theories and are related through a theorem of type Hurewicz from the Moderately Discontinuous Homotopy groups to the MD Homology ones.

With this metric algebraic topology we aim to provide algebraic and numerical invariants which capture Lipschitz phenomena of real and complex analytic singularities, and more generally, subanalytic sets, categories that we hope to enlarge in the future.

The object of study are pointed subanalytic germs, which have conical structure, endowed with a metric that has to be equivalent to a subanalytic one (such as the outer or the inner). A base point in a germ (X,0)(X,0) is what we call a point in these theories, that is a mapping from the interval ([0,1],0)([0,1],0) to the germ (X,O)(X,O) preserving up to a constant the distance to the origin.

An (n,b)(n,b)-loop in a metric germ (X,0,d)(X,0,d) is given by a possibly discontinuous map from the cone of the n-cube InI^{n} to (X,O)(X,O) where we allow discontinuities that measured with respect to the metric dd are no bigger than tbt^{b} where tt is the distance to the origin of the germ. This type of mappings are new and we call them weak bb-maps (see Definition 10).

Then, for every b(0,+)b\in(0,+\infty) we consider the set of (n,b)(n,b)-loops up to weak bb-homotopies (see definitions 18), which admit also discontinuities up to order tbt^{b}. This set with the operation of concatenation is a group that we denote by MDπnb(X,d)MD\pi_{n}^{b}(X,d), which is abelian for n2n\geq 2. Moreover there are homomorphisms relating the (n,b)(n,b)-MD Homotopy group with the (n,b)(n,b^{\prime})-MD Homotopy group. The bb-MD homotopy groups of a metric germ (X,O,d)(X,O,d) capture the homotopy nature of the germ up to gaps that measured in the metric dd are of size smaller than tbt^{b}.

In the case of the outer and inner metric, we prove that all the groups MDπnbMD\pi_{n}^{b} are finitely generated abelian groups. Moreover, only finitely many homomorphisms MDπbMDπbMD\pi^{b}_{\bullet}\to MD\pi^{b^{\prime}}_{\bullet} are essential.

The MD homotopy groups are a bi-Lipschitz subanalytic invariant. They are functorial from the category of metric subanalytic germs, both with lipschitz maps and bb-maps (which are a kind of piecewise continuous mappings that were introduced in [8], see Definition 14) as morphism. These groups are also invariant by suitable metric homotopies, and satisfy versions of the relative long exact sequence and they are independent of the base point for a bb-connected germ (in the sense of [8], see Definition 38).

We prove a Seifert van Kampen type theorem for the bb-MD fundamental groupoid for coverings good enough. That is, we get conditions so that it is the colimit of the bb-MD fundamental groupoids of the elements of the covering as in the classical theorem (see Theorem 55). Seifert van Kampen Theorem is one of the delicate aspects of our theory. This is no surprise, since Mayer-Vietoris for MDMD-Homology was also delicate to formulate and prove. One subtlety lies in finding the appropriate notion of covering for which the result may hold. There are two conditions (()b(*)_{b} and ()b(**)_{b}) that have to be satisfied in order that Theorem 55 is satisfied. Before we prove it we show a more general theorem (see Theorem 54) that is satisfied only assuming condition ()b(*)_{b}.

We prove comparison theorems, which relate the MD-homotopy groups of a germ with the outer metric, with usual homotopy group of limits of bb-horn neighbourhood of it (see Theorem 65). This result has an analogous in the MD Homology theory that was conjetured by Lev Birbrair and proved in [8]. Here it is not possible to adapt the proof of the corresponding result for homology, since it is based in Mayer-Vietoris sequences, that are not available in homotopy. Instead, we need to perform an interpolation procedure that ”metrically homotopes” a discontinuous weak bb-map to a continuous one (see the proof of Proposition 63 and its preparations for details.).

The comparison theorems have many important consequences for pointed pairs of metric germs with the inner or the outer metric: the Hurewicz isomorphims theorem is satisfied (see Theorem 47 and its proof at the end of Section 5.2), the MD homotopy groups are finitely presented, given a pointed pair of metric subanalytic sets, the set of bb’s such that the groups MDπkb+ϵMD\pi^{b+\epsilon}_{k} and MDπkbϵMD\pi^{b-\epsilon}_{k} differ for sufficiently small positive ϵ\epsilon is finite and contained in {\mathbb{Q}}, for b=1b=1 and the outer metric MDπk1MD\pi^{1}_{k} recovers the kk-th homotopy group of the punctured tangent cone, for the inner metric MDπk1MD\pi^{1}_{k} recovers the kk-th homotopy group of the punctured Gromov tangent cone, and MDπkMD\pi^{\infty}_{k} recovers the kk-th homotopy group of the link (see Proposition 58 and Corollary 67). The b=b=\infty comparison theorem (Proposition 58) is easier and has a direct proof. Let us also comment that for the proof of this statements for the inner metric we need to use an adequate re-embedding of the germ that reduces the assertions to the outer metric case.

As a first example we compute the MDMD-homotopy for the bb-cones, completing the computation also for the MDMD-homology started in [8].

The inner metric of normal complex surface singularities is fully described in [3]. In the last section we use this description and the metric version of the Seifert van Kampen theorem, Theorem 55, to compute all the bb-MD fundamental groups of a complex singularity surface germ with the inner metric. Given a surface germ (X,0)(X,0) we give a bb-homotopy model XϵbX^{b}_{\epsilon} which is a topological space whose homology and fundamental group is the bb-MD Homology and bb-MD fundamental group of (X,0,dinn)(X,0,d_{inn}) (see Theorems 78-79). The space XϵbX^{b}_{\epsilon} has the homotopy type of a plumbed 33-manifold in which several circles are identified (a “branched 33-manifold” in the language of [3]). Such a branched 33-manifold has a natural description compatible with a JSJ decomposition of the link XϵX_{\epsilon}.

It is well known that the fundamental group of the link of a normal surface singularity determines the topology of the singularity, except in the cyclic quotient case. This is a particular case of a theorem of Waldhausen. We end the paper with a list of open questions (See Problems 8081), out of which the following two stand out:

  • Does the MD fundamental group of a normal complex surface singularity determine the inner geometry of the surface, in the non cyclic quotient case?

  • Can the Lipschitz normally embedded property of a normal surface singularity be read from all the bb-MD fundamental groups for the inner and the outer metric?

2. Setting and notation

We recall the basic definitions that were introduced in [8], where the reader can find a more detailed exposition. We will always work with bounded subanalytic subsets, which in particular are globally subanalytic (see [6]). Recall that the collection of all globally subanalytic sets forms an O-minimal structure (see [6]). We use [7] and [5] as basic references in O-minimal geometry.

Definition 1.

A metric subanalytic germ (X,x0,d)(X,x_{0},d) is a subanalytic germ (X,x0)(X,x_{0}) such that x0X¯x_{0}\in\overline{X} (where X¯\overline{X} denotes the closure of XX in m{\mathbb{R}}^{m}) and dd is a subanalytic metric that induces the same topology on XX as the restriction of the standard topology on n{\mathbb{R}}^{n} does . We omit x0x_{0} and dd in the notation when it is clear from the context. We say x0x_{0} is the vertex of the germ.

Given two germs (X,x0)(X,x_{0}) and (Y,y0)(Y,y_{0}), a subanalytic map germ f:(X,x0)(Y,y0)f:(X,x_{0})\to(Y,y_{0}) is a subanalytic continuous map f:XYf:X\to Y that admits a continuous and subanalytic extension to a map germ f¯:(X{x0},x0)(Y{y0},y0)\overline{f}:(X\cup\{x_{0}\},x_{0})\to(Y\cup\{y_{0}\},y_{0}).

We define in the expected manner metric subanaytic subgerms, pairs of metric subanalytic germs and subanalytic mappings between them, etc. For more details, see section 2.2 in [8].

We recall that, as in [8], it is possible for a germ (X,x0)(X,x_{0}), according to our definition, that x0Xx_{0}\notin X. Important examples of germs of these types are open subgerms (U,0)(U,0) of a subanalytic germ (X,0)(X,0).

Also, notice that we are building a bilipschitz invariant, so, in practice metrics that are not subanalytic but are bilipchitz equivalent to subanalytic ones are allowed; for example the inner metric. See [8] for more details.

Definition 2.

A map germ f:(X,x0)(Y,y0)f:(X,x_{0})\rightarrow(Y,y_{0}) is said to be linearly vertex approaching (l.v.a. for brevity) if there exists K1K\geq 1 such that

1Kxx0f(x)y0Kxx0\frac{1}{K}||x-x_{0}||\leq||f(x)-y_{0}||\leq K||x-x_{0}||

for every xx in some representative of (X,x0)(X,x_{0}). The constant KK is called the l.v.a constant for ff.

Let (X,x0,dX)(X,x_{0},d_{X}) and (Y,y0,dY)(Y,y_{0},d_{Y}) be two metric subanalytic germs. A Lipschitz linearly vertex approaching subanalytic map germ f:(X,x0,dX)(Y,y0,dY)f:(X,x_{0},d_{X})\rightarrow(Y,y_{0},d_{Y}) is a map germ that is both Lipschitz (with respect to the metrics dXd_{X} and dYd_{Y}) and l.v.a.. In particular one can take the same constant KK for both properties.

Following the philosophy of [8], the metric germ ((0,ϵ),0,d)((0,\epsilon),0,d), that is the interval germ with the euclidean metric, plays the role of a point. This motivates the following definition, which fixes the category of pointed pairs of metric subanalytic germs (the source category of the Lipschitz homotopy functors introduced in this paper).

Definition 3.

The category of pairs of metric subanalytic germs has pairs of metric subanalytic germs as objects and Lipschitz l.v.a. subanalytic maps of pairs as morphisms.

A point in XX where (X,x0,d)(X,x_{0},d) is a metric subanalytic germ, is a continuous l.v.a. subanalytic map germ 𝔭:(0,ϵ)X\mathfrak{p}:(0,\epsilon)\to X. A pointed pair of metric subanalytic germs is a pair (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) of metric subanalytic germs together with a point 𝔭\mathfrak{p} in YY. Given pointed metric subanalytic germs (Y,x0,d,𝔭)(Y,x_{0},d,\mathfrak{p}) and (Y,x0,d,𝔭)(Y^{\prime},x^{\prime}_{0},d^{\prime},\mathfrak{p}^{\prime}), a morphism f:(Y,x0,d)(Y,x0,d)f:(Y,x_{0},d)\to(Y^{\prime},x^{\prime}_{0},d^{\prime}) preserves the base point if f𝔭=𝔭f{\circ}\mathfrak{p}=\mathfrak{p}^{\prime}. The category of pointed pairs of metric subanalytic germs has pointed pairs of metric subanalytic germs as objects and Lipschitz l.v.a. subanalytic maps of pairs that preserves the base point as morphisms.

As in [8] we can enlarge the category admitting morphisms that are bb-maps, see Definition 14.

The following definition is important:

Definition 4.

[bb-point] Let (X,x0,d)(X,x_{0},d) be a metric subanalytic germ and let b(0,+)b\in(0,+\infty). Two points 𝔭\mathfrak{p} and 𝔮\mathfrak{q} in XX are bb-equivalent, and we write 𝔭b𝔮\mathfrak{p}\sim_{b}\mathfrak{q}, if

limt0dX(𝔭(t),𝔮(t))tb=0.\lim_{t\to 0}\frac{d_{X}(\mathfrak{p}(t),\mathfrak{q}(t))}{t^{b}}=0.

An equivalence class of points is called a bb-point of X.

2.1. Useful terminology

In the development of metric homotopy theory we will use nn-loops and homotopies modeled on cones over cubes. We introduce some related notation here:

Notation 5.

In this chapter I:=[0,1]I:=[0,1] denotes the unit interval.

We consider the cone over the cube InI^{n}:

C(In)={(tx,t)n×;xIn,t(0,+)}.C(I^{n})=\{(tx,t)\in{\mathbb{R}}^{n}\times{\mathbb{R}};\,x\in I^{n},\ t\in(0,+\infty)\}.

For convenience the origin is not in C(In)C(I^{n}).

Note that the notion of l.v.a. maps from or into C(In)C(I^{n}) can be thought using the parameter tt instead of the distance to the origin.

We sometimes denote (y1,,yn)In(y_{1},...,y_{n})\in I^{n} by y1..ny_{1..n} or also by (y1..n1,yn)(y_{1..n-1},y_{n}) and similarly.

In a similar way we denote by C(M)C(M) the cone over a bounded subanalytic set MnM\subset{\mathbb{R}}^{n}.

For readability, we write (y,t)(y,t) to denote (yt,t)(yt,t) in C(M)C(M). But be aware that this does only provide a system of coordinates out of the origin of C(M)C(M).

Definition 6 (Normal point in C(In)C(I^{n})).

Let 𝔮:(0,ϵ)C(In)\mathfrak{q}:(0,\epsilon)\to C(I^{n}) be a point in C(In)C(I^{n}). We say that 𝔮\mathfrak{q} is a normal point if 𝔮(t)\mathfrak{q}(t) is expressed as (α(t),t)(\alpha(t),t) in the usual coordinates of C(In)C(I^{n}).

Notation 7.

We respectively denote by 𝒮{\mathcal{S}}, 𝒢{\mathcal{G}}, and 𝒢𝒫{\mathcal{G}}{\mathcal{P}} the categories of sets, groups, and groupoids.

We denote by 𝔹{\mathbb{B}} the category whose set of objects is (0,](0,\infty] and there is a unique morphism from bb to bb^{\prime} if and only if bbb\geq b^{\prime}.

The category 𝔹𝒮{\mathbb{B}}-{\mathcal{S}}, (resp. 𝔹𝒢{\mathbb{B}}-{\mathcal{G}}, 𝔹𝒢𝒫{\mathbb{B}}-{\mathcal{G}}{\mathcal{P}}) of 𝔹{\mathbb{B}}-sets (resp. 𝔹{\mathbb{B}}-groups, 𝔹{\mathbb{B}}-groupoids) is the category whose objects are functors from 𝔹{\mathbb{B}} to 𝒮{\mathcal{S}} (resp. 𝒢{\mathcal{G}}, 𝒢𝒫{\mathcal{G}}{\mathcal{P}}) and the morphisms are natural transformations of functors.

2.2. Conical structures

Given a subanalytic germ (X,x0)(X,x_{0}), its link CXC_{X} is the intersection of XX with a small sphere centered in x0x_{0}. Its subanalytic homeomorphism type is independent of the radius.

Definition 8.

Given a subanalytic germ (X,x0)(X,x_{0}) and a family of subanalytic subgerms (Z1,0)(Z_{1},0),…,(Zk,0)(X,0)(Z_{k},0)\subseteq(X,0), a conical structure for (X,0)(X,0) compatible with the the family {Zi}\{Z_{i}\} is a subanalytic homeomorphism h:C(LX)(X,x0)h:C(L_{X})\to(X,x_{0}) such that x0h(x,t)=t||x_{0}-h(x,t)||=t and such that h(C(LZi))=Zih(C(L_{Z_{i}}))=Z_{i} with LZiL_{Z_{i}} in LXL_{X}.

Conical structures always exist (see [5] Theorem 4.10). The following remark, proved in [8] will be used.

Remark 9.

Let (X,x0)(X,x_{0}) be a subanalytic germ with compact link. Consider any subanalytic map germ f:(X,x0)(Y,y0)f:(X,x_{0})\rightarrow(Y,y_{0}) that is a homeomorphism onto its image. Let {Zj}jJ\{Z_{j}\}_{j\in J} be a finite collection of closed subanalytic subsets of XX. There is a subanalytic homeomorphism germ ϕ:(X,x0)(X,x0)\phi:(X,x_{0})\to(X,x_{0}) such that ϕ(Zj)=Zj\phi(Z_{j})=Z_{j} for all jJj\in J and such that fϕ(x)y0=xx0||f{\circ}\phi(x)-y_{0}||=||x-x_{0}|| (this is stronger than l.v.a.).

3. bb-moderately discontinuous homotopy groups and their basic properties

3.1. Weak bb-maps

Weak bb-maps are a way of weakening the continuity of loops and homotopies in the classical theory, in order to establish a parallel theory that captures metric phenomena.

Definition 10 (Weak bb-map).

Let (X,x0,dX)(X,x_{0},d_{X}) be a metric subanalytic germ. Let b(0,)b\in(0,\infty). A weak bb-moderately discontinuous subanalytic map (weak bb-map, for abbreviation) from a subanalytic germ (Z,0)(Z,0) to (X,x0,d)(X,x_{0},d) is a finite collection {(Cj,fj)}jJ\{(C_{j},f_{j})\}_{j\in J}, where {Cj}jJ\{C_{j}\}_{j\in J} is a finite closed subanalytic cover of (Z,0¯)(Z,\underline{0}) and fj:CjXf_{j}:C_{j}\to X are continuous l.v.a. subanalytic maps for which for any point 𝔮\mathfrak{q} in Cj1Cj2C_{j_{1}}\cap C_{j_{2}} for any j1,j2Jj_{1},j_{2}\in J we have that fj1𝔮bfj2𝔮f_{j_{1}}{\circ}\mathfrak{q}\sim_{b}f_{j_{2}}{\circ}\mathfrak{q}. We call {Cj}jJ\{C_{j}\}_{j\in J} the cover of the weak bb-map {(Cj,fj)}jJ\{(C_{j},f_{j})\}_{j\in J}.

Two weak bb-maps {(Cj,fj)}jJ\{(C_{j},f_{j})\}_{j\in J} and {(Ck,fk)}kK\{(C^{\prime}_{k},f^{\prime}_{k})\}_{k\in K} from (Z,0)(Z,0) to (X,x0,d)(X,x_{0},d) are called bb-equivalent, if for any point 𝔮\mathfrak{q} contained in the intersection CjCkC_{j}\cap C^{\prime}_{k} for any jJj\in J and kKk\in K, we have that fj𝔮bfk𝔮f_{j}\circ\mathfrak{q}\sim_{b}f^{\prime}_{k}{\circ}\mathfrak{q}.

We make an abuse of language and we also say that a weak bb-map from (Z,z0)(Z,z_{0}) to (X,x0,dX)(X,x_{0},d_{X}) is an equivalence class as above.

For b=b=\infty, a weak bb-map from ZZ to XX is a continuous l.v.a. subanalytic map germ from (Z,z0)(Z,z_{0}) to (X,x0,dX)(X,x_{0},d_{X}).

Informally we can say that a weak bb-map gives a well defined mapping from ZZ to the set of bb-points XX (see Definition 4).

Remark 11 (Gluing of weak bb-maps).

Two weak bb-maps φ1\varphi_{1} and φ2\varphi_{2} defined on Z1Z_{1} and Z2Z_{2} respectively glue to a weak bb-map defined on Z1Z2Z_{1}\cup Z_{2} if and only if for any point 𝔮\mathfrak{q} in Z1Z2Z_{1}\cap Z_{2} we have the equivalence φ1𝔮bφ2𝔮\varphi_{1}{\circ}\mathfrak{q}\sim_{b}\varphi_{2}{\circ}\mathfrak{q}.

Remark 12 (Equivalence by refinement).

Let φ={(Cj,fj)}jJ\varphi=\{(C_{j},f_{j})\}_{j\in J} be a weak bb-map and {Dk}kK\{D_{k}\}_{k\in K} a refinement of {Cj}jJ\{C_{j}\}_{j\in J}. For kKk\in K, let r(k)Jr(k)\in J be such that DkCr(k)D_{k}\subseteq C_{r(k)}. Then {(Dk,fr(k)|Dk)}kK\{(D_{k},f_{r(k)}|_{D_{k}})\}_{k\in K} is equivalent to φ\varphi. As a consequence any weak bb-map from ZZ to XX has a representative {(Cj,fj)}jJ\{(C_{j},f_{j})\}_{j\in J}, for which the interior of Cj1Cj2C_{j_{1}}\cap C_{j_{2}} is empty for any j1,j2Jj_{1},j_{2}\in J.

Remark 13.

If bbb\geq b^{\prime} then any weak bb-map is also a weak bb^{\prime}-map.

We recall the definition of bb-maps, with respect to which the bb-MD-Homology groups are functorial (see Section 5 in [8]). We will also prove functoriality of the MD-Homotopy groups for this type of morphisms. In particular, we can compose a weak bb-map with a bb-map on the right, as we see in Definition 15 below.

Definition 14 (Category pointed of metric pairs with bb-maps).

Let (X,x0,dX)(X,x_{0},d_{X}) and (Y,y0,dY)(Y,y_{0},d_{Y}) be metric subanalytic germs, b(0,+)b\in(0,+\infty). A bb-moderately discontinuous subanalytic map (a bb-map, for abbreviation) from (X,x0,dX)(X,x_{0},d_{X}) to (Y,y0,dY)(Y,y_{0},d_{Y}) is a finite collection {(Ci,fi)}iI\{(C_{i},f_{i})\}_{i\in I}, where {Ci}iI\{C_{i}\}_{i\in I} is a finite closed subanalytic cover of XX and the fi:CiYf_{i}:C_{i}\to Y are l.v.a. subanalytic maps satisfying the following: for any bb-equivalent pair of points 𝔭\mathfrak{p} and 𝔮\mathfrak{q} contained in CiC_{i} and CjC_{j} respectively (ii and jj may be equal), the points fi𝔭f_{i}\circ\mathfrak{p} and fj𝔮f_{j}\circ\mathfrak{q} are b-equivalent in YY.

Two bb-maps {(Ci,fi)}iI\{(C_{i},f_{i})\}_{i\in I} and {(Ci,fi)}iI\{(C^{\prime}_{i},f^{\prime}_{i})\}_{i\in I^{\prime}} are called bb-equivalent if for any bb-equivalent pair of points 𝔭\mathfrak{p}, 𝔮\mathfrak{q} with Im(𝔭)Ci(\mathfrak{p})\subseteq C_{i} and Im(𝔮)Ci(\mathfrak{q})\subseteq C^{\prime}_{i^{\prime}}, the points fi𝔭f_{i}\circ\mathfrak{p} and fi𝔮f^{\prime}_{i^{\prime}}\circ\mathfrak{q} are b-equivalent in YY.

We make an abuse of language and we also say that a bb-map from (X,x0,dX)(X,x_{0},d_{X}) to (Y,y0,dY)(Y,y_{0},d_{Y}) is an equivalence class as above.

For b=b=\infty, a bb-map from XX to YY is a l.v.a. subanalytic map from XX to YY.

A bb-map between pointed pairs of metric subanalytic germs (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) and (X~,Y~,x~0,d~,𝔭~)(\widetilde{X},\widetilde{Y},\widetilde{x}_{0},\widetilde{d},\widetilde{\mathfrak{p}}) is a bb-map from XX to X~\widetilde{X} admitting a representative {(Ci,fi)}iI\{(C_{i},f_{i})\}_{i\in I} for which

  1. (1)

    for any point 𝔮\mathfrak{q} of YCiY\cap C_{i}, the point fi𝔭f_{i}{\circ}\mathfrak{p} is bb-equivalent to a point in Y~\widetilde{Y},

  2. (2)

    if the image of 𝔭\mathfrak{p} is in CiC_{i} then fi𝔭b𝔭~f_{i}{\circ}\mathfrak{p}\sim_{b}\widetilde{\mathfrak{p}}.

The category of pointed metric pairs with bb-maps has as objects pointed pairs of metric subanalytic germs and as morphisms the bb-maps between them.

Informally, we can say that bb-maps give well defined mappings from the set of bb-points of XX to the set of bb-points of YY.

Note that the analogue for bb-maps of Remarks 11 and 13 for weak bb-maps are not satisifed (unless the target of the bb-map is a convex set; see [10] for more details).

Definition 15 (Composition of weak bb-maps and bb-maps).

Let ZZ and ZZ^{\prime} be subanalytic germs and let (X,x0,d)(X,x_{0},d) and (X,x0,d)(X^{\prime},x_{0},d^{\prime}) be metric subanalytic germs. Let φ={(Cj,fj)}jJ\varphi=\{(C_{j},f_{j})\}_{j\in J} be a weak bb-map from ZZ to XX.

For a continuous l.v.a. subanalytic map ϕ\phi from ZZ^{\prime} to ZZ, we can define φϕ\varphi{\circ}\phi to be the weak bb-map {(ϕ1(Cj),φjϕ)}jJ\{(\phi^{-1}(C_{j}),\varphi_{j}{\circ}\phi)\}_{j\in J} from ZZ^{\prime} to XX.

Let ψ=(Dk,gk)kK\psi=(D_{k},g_{k})_{k\in K} be a bb-map from XX to XX^{\prime}. We define ψφ\psi{\circ}\varphi to be the weak bb-map {(fj1(Dk)Cj,gkfj|fj1(Dk)Cj)}(j,k)J×K\{(f_{j}^{-1}(D_{k})\cap C_{j},g_{k}\circ f_{j|f_{j}^{-1}(D_{k})\cap C_{j}})\}_{(j,k)\in J\times K} from ZZ to XX^{\prime}.

3.2. Definition of the bb-moderately discontinuous metric homotopy groups

Definition 16 ((n,b)(n,b)-loop).

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) be a pointed pair of metric subanalytic germs, nn\in\mathbb{N} and b(0,]b\in(0,\infty]. A bb-moderately discontinuous nn-loop (a ((n,b)(n,b)-loop, for short) is a weak bb-map φ\varphi from C(In)C(I^{n}) to (X,Y)(X,Y) for which the following boundary conditions hold:

  1. (a)

    for any point 𝔮\mathfrak{q} in C(In)C(\partial I^{n}), the point φ𝔮\varphi{\circ}\mathfrak{q} is bb-equivalent to a point in YY.

  2. (b)

    for any normal point 𝔮\mathfrak{q} in C(In(In1×{1}))C(\partial I^{n}\setminus(I^{n-1}\times\{1\})), we have φ𝔮b𝔭\varphi{\circ}\mathfrak{q}\sim_{b}\mathfrak{p}.

We denote the set of all (n,b)(n,b)-loops in (X,Y,𝔭)(X,Y,\mathfrak{p}) by MDΓnb(X,Y,𝔭)MD\Gamma_{n}^{b}(X,Y,\mathfrak{p}). Observe that we suppress x0x_{0} and dd in the notation MDΓnb(X,Y,𝔭)MD\Gamma_{n}^{b}(X,Y,\mathfrak{p}), even though they influence the set of (n,b)(n,b)-loops in (X,Y,𝔭)(X,Y,\mathfrak{p}). In the case YY coincides with the image of 𝔭\mathfrak{p}, we simply write MDΓnb(X,𝔭)MD\Gamma_{n}^{b}(X,\mathfrak{p})

Notation 17.

Denote the inclusions ιs:C(In)C(In+1)\iota_{s}:C(I^{n})\to C(I^{n+1}) defined as ιs(y,t):=((y,s),t)\iota_{s}(y,t):=((y,s),t) for any s[0,1]s\in[0,1], and the projection ρ:C(In+1)C(In)\rho:C(I^{n+1})\to C(I^{n}) defined by ρ(y1..n+1,t):=(y1..n,t)\rho(y_{1..n+1},t):=(y_{1..n},t).

Definition 18 (Weak bb-homotopy (relative to WW)).

Let (X,x0,d)(X,x_{0},d) be a metric subanalytic germ. Let φ0\varphi_{0}, φ1\varphi_{1} be weak bb-maps from C(In)C(I^{n}) to XX . A weak bb-homotopy from φ0\varphi_{0} to φ1\varphi_{1} is a weak bb-map HH from C(In+1)C(I^{n+1}) to XX such that Hι0=φ0H\circ\iota_{0}=\varphi_{0} and Hι1=φ1H\circ\iota_{1}=\varphi_{1} where ιs\iota_{s} denotes the inclusion of C(In)C(I^{n}) into C(In+1)C(I^{n+1}) given by (y,t)((y,s),t)(y,t)\to((y,s),t).

We say HH is a weak bb-homotopy relative to a subgerm (W,0¯)(C(In),0¯)(W,\underline{0})\subseteq(C(I^{n}),\underline{0}) if moreover

H𝔮bφ0ρ𝔮H{\circ}\mathfrak{q}\sim_{b}\varphi_{0}{\circ}\rho{\circ}\mathfrak{q}

for any point 𝔮\mathfrak{q} in ρ1(W)\rho^{-1}(W).

In case φ0\varphi_{0}, φ1\varphi_{1} satisfy that for any point 𝔭\mathfrak{p} in a subanalytic germ KK of C(In)C(I^{n}) the points φ0𝔭\varphi_{0}\circ\mathfrak{p} and φ1𝔭\varphi_{1}\circ\mathfrak{p} are bb-equivalent to points in certain subgerm (Y,x0)(Y,x_{0}) of (X,x0)(X,x_{0}), and for any point 𝔮\mathfrak{q} in ρ1(K)\rho^{-1}(K) the point H𝔮H{\circ}\mathfrak{q} is bb-equivalent to a point in YY then we say that HH preserves the inclusion of KK in YY.

We say that φ0\varphi_{0} and φ1\varphi_{1} are weakly bb-homotopically equivalent or weak bb-homotopic (relative to WW or preserving the inclusion of KK in YY if it applies).

Refer to caption
Figure 1. Squematic example of a weak bb-homotopy joining the weak bb-maps α\alpha and β\beta in blue and purple.
Remark 19.

Let φ0\varphi_{0} and φ1\varphi_{1} be weak bb-maps from C(In)C(I^{n}) to XX. Let (W,0)(C(In),0)(W,0)\subseteq(C(I^{n}),0) be a subgerm. There is a necessary condition for φ0\varphi_{0} and φ1\varphi_{1} to admit a weak bb-homotopy relative to WW between them: for any point 𝔮\mathfrak{q} in WW, the points φ0𝔮\varphi_{0}{\circ}\mathfrak{q} and φ1𝔮\varphi_{1}{\circ}\mathfrak{q} are bb-equivalent.

In particular any two (n,b)(n,b)-loops φ1\varphi_{1} and φ2\varphi_{2} in (X,Y,𝔭)(X,Y,\mathfrak{p}) fulfill the necessary condition to admit a weak bb-homotopy relative to W=C(InIn1×{1})W=C(\partial I^{n}\setminus I^{n-1}\times\{1\}) between them. Moreover, if HH is such weak bb-homotopy, then Hιs=φsH{\circ}\iota_{s}=\varphi_{s} is a bb-moderately discontinuous nn-loop for any s[0,1]s\in[0,1].

Proof.

We prove the last statement. Let 𝔮\mathfrak{q} be a point in C(InIn1×{1})C(\partial I^{n}\setminus I^{n-1}\times\{1\}). By Remark 9, there is a subanalytic homeomorphism h:(0,ϵ)(0,ϵ)h:(0,\epsilon)\to(0,\epsilon), for which 𝔮~:=𝔮h\widetilde{\mathfrak{q}}:=\mathfrak{q}{\circ}h is a normal point. Since both 𝔮\mathfrak{q} and 𝔮~\widetilde{\mathfrak{q}} are l.v.a., the homeomorphism hh is also l.v.a.. Therefore, the bb-equivalence between φ1q~\varphi_{1}{\circ}\widetilde{q} and φ2q~\varphi_{2}{\circ}\widetilde{q} implies the bb-equivalence between φ1𝔮\varphi_{1}{\circ}\mathfrak{q} and φ2𝔮\varphi_{2}{\circ}\mathfrak{q}. ∎

Let us see some easy ways of getting weak bb-homotopies between weak bb-maps:

Example 20.

Let η:C(In)×IC(In)\eta:C(I^{n})\times I\to C(I^{n}) be a subanalytic continuous homotopy with η0=idC(In)\eta_{0}=id_{C(I^{n})} and satisfying that there is a K1K\geq 1 such that for any sIs\in I, ηs:=η(_,s)\eta_{s}:=\eta(\_,s) is l.v.a. for the constant KK. We define η^:C(In+1)C(In)\hat{\eta}:C(I^{n+1})\to C(I^{n}) by the formula η^(y1..n+1,t):=η((y1..n,t),yn+1)\hat{\eta}(y_{1..n+1},t):=\eta((y_{1..n},t),y_{n+1}). Then φη^\varphi{\circ}\hat{\eta} defines a weak bb-homotopy from φ\varphi to φη1\varphi{\circ}\eta_{1}.

We are going to define concatenations of weak bb-maps and equip MDπnb(X,𝔭)MD\pi_{n}^{b}(X,\mathfrak{p}) with a product operation, endowing it with a group structure in most cases.

A concatenation of (n,b)(n,b)-loops φ0\varphi_{0} and φ1\varphi_{1} is defined, similarly to the classical case, by gluing them along the faces C({1}×In1)C(\{1\}\times I^{n-1}) and C({0}×In1)C(\{0\}\times I^{n-1}).

We will define concatenation for weak bb-maps that are not necessarily (n,b)(n,b)-loops. The following auxiliary mappings will be used:

Notation 21.

Let nn\in\mathbb{N} and let 0α1α210\leq\alpha_{1}\leq\alpha_{2}\leq 1 and 0α1<α210\leq\alpha^{\prime}_{1}<\alpha^{\prime}_{2}\leq 1. Then, ϕα1,α2α1,α2\phi_{\alpha_{1},\alpha_{2}}^{\alpha^{\prime}_{1},\alpha^{\prime}_{2}} denotes the continuous subanalytic l.v.a. homeomorphism from C([α1,α2]×In1)C([\alpha^{\prime}_{1},\alpha^{\prime}_{2}]\times I^{n-1}) to C([α1,α2]×In1)C([\alpha_{1},\alpha_{2}]\times I^{n-1}) that linearly transforms the former into the latter. This is defined by the formula

ϕα1,α2α1,α2(y1..n,t):=((α2α2α1α2α1(α2y1),y2..n),t)\phi_{\alpha_{1},\alpha_{2}}^{\alpha^{\prime}_{1},\alpha^{\prime}_{2}}(y_{1..n},t):=((\alpha_{2}-\frac{\alpha_{2}-\alpha_{1}}{\alpha^{\prime}_{2}-\alpha^{\prime}_{1}}(\alpha^{\prime}_{2}-y_{1}),y_{2..n}),t)

We suppress nn in the notation. When α1=0\alpha^{\prime}_{1}=0 and α2=1\alpha^{\prime}_{2}=1, we simply write ϕα1,α2\phi_{\alpha_{1},\alpha_{2}}.

Remark 22.

Let nn\in\mathbb{N}, 0α1<α210\leq\alpha_{1}<\alpha_{2}\leq 1 and 0β1<β210\leq\beta_{1}<\beta_{2}\leq 1. Then we have ϕβ1,β2ϕα1,α2=ϕγ1,γ2\phi_{\beta_{1},\beta_{2}}{\circ}\phi_{\alpha_{1},\alpha_{2}}=\phi_{\gamma_{1},\gamma_{2}}, where γ1=α1(β2β1)+β1\gamma_{1}=\alpha_{1}(\beta_{2}-\beta_{1})+\beta_{1} and γ2=α2(β2β1)+β1\gamma_{2}=\alpha_{2}(\beta_{2}-\beta_{1})+\beta_{1}.

Definition 23 (Concatenation).

Let (X,x0,d)(X,x_{0},d) be a metric subanalytic germ. Let φ0\varphi_{0} and φ1\varphi_{1} be weak bb-maps from C(In)C(I^{n}) to XX. Assume that for any point 𝔮\mathfrak{q} in C(In1)C(I^{n-1}) we have that

(1) φ0ν1𝔮bφ1ν0𝔮\varphi_{0}\circ\nu_{1}\circ\mathfrak{q}\sim_{b}\varphi_{1}\circ\nu_{0}\circ\mathfrak{q}

where νs:C(In1)C(In)\nu_{s}:C(I^{n-1})\to C(I^{n}) are the inclussion defined by νs(y,t)=(s,y,t)\nu_{s}(y,t)=(s,y,t). By Remark 11, φ0(ϕ0,12)1\varphi_{0}{\circ}(\phi_{0,\frac{1}{2}})^{-1} and φ1(ϕ12,1)1\varphi_{1}{\circ}(\phi_{\frac{1}{2},1})^{-1} glue to a weak bb-map on C(In)C(I^{n}), which we call the concatenation of φ0\varphi_{0} and φ1\varphi_{1}. We denote it by φ0φ1\varphi_{0}\cdot\varphi_{1}.

Then, one can always concatenate (n,b)(n,b)-loops in MDΓnb(X,Y,𝔭)MD\Gamma_{n}^{b}(X,Y,\mathfrak{p}) if n>1n>1, and in MDΓnb(X,𝔭)MD\Gamma_{n}^{b}(X,\mathfrak{p}) if n1n\geq 1. Since weak bb-homotopies between (n,b)(n,b)-loops can be similarly concatenated, we get that concatenation is well defined up to weak bb-homotopies preserving the inclusions of C(In)C(\partial I^{n}) into YY and relative to C(InIn1×{1})C(\partial I^{n}\setminus I^{n-1}\times\{1\}). In particular we have a group structure on the equivalence classes of (n,b)(n,b)-loops up to homotopy whenever concatenation is possible as above. See Proposition 26 for the final statement.

Notation 24.

[Constant loop] Let (X,x0,d,𝔭)(X,x_{0},d,\mathfrak{p}) be a pointed metric subanalytic germ and nn\in\mathbb{N}. We denote by c𝔭,nc_{\mathfrak{p},n} the weak bb-map from C(In)C(I^{n}) to XX defined by c𝔭,n(y,t)=𝔭(t)c_{\mathfrak{p},n}(y,t)=\mathfrak{p}(t).

Notation 25.

[Inverse loop] Let φ={(Cj,fj)}\varphi=\{(C_{j},f_{j})\} be a weak bb-map from C(In)C(I^{n}) to XX. We denote the weak bb-map {(Cj,fj)}\{(\overleftarrow{C_{j}},\overleftarrow{f_{j}})\} by φ\overleftarrow{\varphi}, where Cj\overleftarrow{C_{j}} and fj\overleftarrow{f_{j}} are the result of mirroring CjC_{j} and fjf_{j} respectively at the y1y_{1}-axis:

Cj:={(y1..n,t)C(In):(1y1,y2..n,t)Cj},\displaystyle\overleftarrow{C_{j}}:=\{(y_{1..n},t)\in C(I^{n}):(1-y_{1},y_{2..n},t)\in C_{j}\},
fj(y1..n,t):=fj(1y1,y2..n,t)\displaystyle\overleftarrow{f_{j}}(y_{1..n},t):=f_{j}(1-y_{1},y_{2..n},t)

The notation φa\varphi^{-a} stands for the result of concatenating φ\overleftarrow{\varphi} with itself aa times.

Proposition 26 (Definition of the (n,b)-MD homotopy groups MDπnb(X,Y,𝔭)MD\pi_{n}^{b}(X,Y,\mathfrak{p})).

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) be a pointed pair of metric subanalytic germs. We denote by MDπnb(X,Y,𝔭)MD\pi_{n}^{b}(X,Y,\mathfrak{p}) the quotient of MDΓnb(X,Y,𝔭)MD\Gamma_{n}^{b}(X,Y,\mathfrak{p}) by weak bb-homotopies that preserve the inclusion of In\partial I^{n} into YY and are relative to C(In(In1×{1}))C(\partial I^{n}\setminus(I^{n-1}\times\{1\})). We denote by [φ][\varphi] the equivalence class in MDπnb(X,Y,𝔭)MD\pi_{n}^{b}(X,Y,\mathfrak{p}) of an element φMDΓnb(X,Y,𝔭)\varphi\in MD\Gamma_{n}^{b}(X,Y,\mathfrak{p}). We call it the bb-homotopy class of φ\varphi.

For n>1n>1 concatenation of (n,b)(n,b)-loops is possible as defined in Definition 23. It induces a well defined operation of weak bb-homotopy classes as [φ1][φ2]:=[φ1φ2][\varphi_{1}]\cdot[\varphi_{2}]:=[\varphi_{1}\cdot\varphi_{2}] that defines a group structure on MDπnb(X,Y,𝔭)MD\pi_{n}^{b}(X,Y,\mathfrak{p}). We call it the (n,b)(n,b)-homotopy group of (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}).

For n=1n=1, concatenation of (1,b)(1,b)-loops in MDΓ1b(X,𝔭)MD\Gamma_{1}^{b}(X,\mathfrak{p}) is always possible and it induces a group structure in MDπ1b(X,𝔭)MD\pi_{1}^{b}(X,\mathfrak{p}). We call MDπ1b(X,𝔭)MD\pi_{1}^{b}(X,\mathfrak{p}) the bb-MD fundamental group of (X,𝔭)(X,\mathfrak{p}).

In these groups, the neutral element is [cp,n][c_{p,n}] and the inverse is [φ]1=[φ][\varphi]^{-1}=[\overleftarrow{\varphi}].

If n3n\geq 3 the group MDπnb(X,Y,𝔭)MD\pi_{n}^{b}(X,Y,\mathfrak{p}) is abelian. If n2n\geq 2 the group MDπnb(X,𝔭)MD\pi_{n}^{b}(X,\mathfrak{p}) is abelian.

Proof.

The proof consist in a routine checking that the corresponding homotopies can be concatenated and the usual arguments in the topological case extended to weak bb-maps. See [10] for details. ∎

The proof of the equality [φ]1=[φ][\varphi]^{-1}=[\overleftarrow{\varphi}] yields the following slightly more general statement that will be used:

Lemma 27.

Let φ={(Cj,fj)}jJ\varphi=\{(C_{j},f_{j})\}_{j\in J} be a weak bb-map from C(I)C(I) to XX. Consider 𝔮0:=φ|C{0}\mathfrak{q}_{0}:=\varphi|_{C\{0\}} and 𝔮1:=φ|C{1}\mathfrak{q}_{1}:=\varphi|_{C\{1\}} and the associated constant loops C𝔮0,1C_{\mathfrak{q}_{0},1} and C𝔮1,1C_{\mathfrak{q}_{1},1}. Then there is a weak bb-homotopy relative to C(I)C(\partial I) from φφ\varphi\cdot\overleftarrow{\varphi} to C𝔮0,1C_{\mathfrak{q}_{0},1} and from φφ\overleftarrow{\varphi}\cdot\varphi to C𝔮1,1C_{\mathfrak{q}_{1},1}.

3.3. Functoriality

Proposition 28.

Let nn\in\mathbb{N} and b(0,]b\in(0,\infty]. There are functorial assignments (X,Y,x0,d,𝔭)MDπ1b(X,Y,𝔭)(X,Y,x_{0},d,\mathfrak{p})\mapsto MD\pi^{b}_{1}(X,Y,\mathfrak{p}) from the category of pointed pairs of metric subanalytic germs with bb-maps to the category 𝒮{\mathcal{S}} of sets. The assignment takes place in the categories 𝒢{\mathcal{G}} or 𝒜𝒢{\mathcal{A}}{\mathcal{G}} of groups or abelian groups when the product is defined.

Proof.

This is a corollary of the fact that weak bb-maps can be composed with bb-maps (see Definition 15). ∎

As we did in the Moderately Discontinuous Homology groups we can enrich the invariant of the MD homotopy groups giving them the structure of an object in the category 𝔹𝒮{\mathbb{B}}-{\mathcal{S}} (recall Notation 7).

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) be a pointed pair of metric subanalytic germs. For any nn\in{\mathbb{N}} and any b,b(0,]b,b^{\prime}\in(0,\infty] with bbb\geq b^{\prime} using the obvious we get a map (which respects the product whenever it is defined)

(2) ηb,b:MDπnb(X,Y,𝔭)MDπnb(X,Y,𝔭)\eta_{b,b^{\prime}}:MD\pi^{b}_{n}(X,Y,\mathfrak{p})\to MD\pi^{b^{\prime}}_{n}(X,Y,\mathfrak{p})
Proposition 29.

Let nn\in\mathbb{N}. There are functorial assignments (X,Y,x0,d,𝔭)MDπ1(X,Y,𝔭)(X,Y,x_{0},d,\mathfrak{p})\mapsto MD\pi^{\star}_{1}(X,Y,\mathfrak{p}) from the category of pointed metric subanalytic germs with Lipschitz subanalytic l.v.a. maps to 𝔹𝒮{\mathbb{B}}-{\mathcal{S}}. The assignment takes place in 𝔹𝒢{\mathbb{B}}-{\mathcal{G}} when the product is defined.

Proof.

This is a consequence of the previous proposition and the fact that Lipschitz l.v.a. maps are bb-maps for any bb. ∎

Notation 30.

Given a bb-map g:(X,𝔭)(X,𝔭)g:(X,\mathfrak{p})\to(X^{\prime},\mathfrak{p}^{\prime}) we denote by gbg^{b}_{\ast} the induced group homomorphisms MDπb(X,𝔭)MDπb(X,𝔭)MD\pi_{\ast}^{b}(X,\mathfrak{p})\to MD\pi_{\ast}^{b}(X^{\prime},\mathfrak{p}^{\prime}) .

3.4. Loops which are small with respect to a dense subgerm

This section is the analogue of Section 4.4 in [8].

Definition 31.

Let (X,Y,d,x0,𝔭)(X,Y,d,x_{0},\mathfrak{p}) be a pointed pair of metric subanalytic germs and let UU be a subanalytic subset of XX. A weak bb-map φ:C(In)X\varphi:C(I^{n})\to X is small with respect to UU if there exists a representative {(Cj,fj}jJ\{(C_{j},f_{j}\}_{j\in J} such that for fj(Cj)Uf_{j}(C_{j})\subset U for every kk. We denote by MDΓkb,U(X,Y,𝔭)MD\Gamma^{b,U}_{k}(X,Y,\mathfrak{p}) the set of (b,k)(b,k)-loops small with respect to UU, and by MDπkb,U(X,Y,𝔭)MD\pi^{b,U}_{k}(X,Y,\mathfrak{p}) the set of equivalence classes of (b,k)(b,k)-loops small with respect to UU, modulo bb-homotopies relative to C(In(Ik1×{1})C(\partial I^{n}\setminus(\partial I^{k-1}\times\{1\}), preserving the inclusion of C(In)C(\partial I^{n}) into YY and which are small with respect to UU.

Proposition 32.

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) be a pointed pair of metric subanalytic germs and let UU be a dense subanalytic subset of XX.

For any b<b<\infty the natural maps

MDΓkb,U(X,Y,𝔭)MDΓkb(X,Y,𝔭)MD\Gamma^{b,U}_{k}(X,Y,\mathfrak{p})\to MD\Gamma^{b}_{k}(X,Y,\mathfrak{p})
MDπkb,U(X,Y,𝔭)MDπkb(X,Y,𝔭)MD\pi^{b,U}_{k}(X,Y,\mathfrak{p})\to MD\pi^{b}_{k}(X,Y,\mathfrak{p})

are bijective for any kk.

Proof.

Injectivity is clear. For surjectivity, a direct application of the Claim in the proof of Proposition 46 in [8] allows to modify the maps fif_{i} within their bb-equivalence class so that fj(Cj)Uf_{j}(C_{j})\subset U as needed. ∎

As in MD-Homology we have the following corollary:

Corollary 33.

Let (X,x0,d)(X,x_{0},d) be a metric subanalytic germ such that the metric dd extends to a subanalytic metric d¯\overline{d} in the closure X¯\overline{X} of XX in n{\mathbb{R}}^{n}. Then for any b<b<\infty we have a bijection

MDπkb(X,x0,𝔭)MDπkb(X¯,x0,𝔭)MD\pi_{k}^{b}(X,x_{0},\mathfrak{p})\cong MD\pi_{k}^{b}(\overline{X},x_{0},\mathfrak{p})

for any kk.

3.5. The long homotopy sequence of a pair

Theorem 34.

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) be a pointed pair of closed metric subanalytic sets. There are functorial assignments:

  • From the category of pointed pairs of metric subanalytic sets with bb-maps to long exact sequences of sets (respecting the product when it is defined)

    MDπkb(Y,𝔭)MDπkb(X,𝔭)MDπkb(X,Y,𝔭)MDπk1b(Y,𝔭)...\to MD\pi^{b}_{k}(Y,\mathfrak{p})\to MD\pi^{b}_{k}(X,\mathfrak{p})\to MD\pi^{b}_{k}(X,Y,\mathfrak{p})\to MD\pi^{b}_{k-1}(Y,\mathfrak{p})\to...
  • From the category of pointed pairs of metric subanalytic sets with Lipschitz subanalytic l.v.a. maps to long exact sequences of 𝔹{\mathbb{B}}-sets (respecting the product when it is defined)

    MDπk(Y,𝔭)MDπk(X,𝔭)MDπk(X,Y,𝔭)MDπk1(Y,𝔭)...\to MD\pi^{\star}_{k}(Y,\mathfrak{p})\to MD\pi^{\star}_{k}(X,\mathfrak{p})\to MD\pi^{\star}_{k}(X,Y,\mathfrak{p})\to MD\pi^{\star}_{k-1}(Y,\mathfrak{p})\to...

The proof has the same steps as in the topological setting, but needs a preliminary result. Let us start recalling Lemma 61 of [8] for the convenience of the reader.

Lemma 35.

Suppose that SQS\supset Q are compact subanalytic subsets in n{\mathbb{R}}^{n}. Let dd be a subanalytic metric in SS. There exists a partition of SS into finitely many disjoint subanalytic subsets {Sk}k\{S_{k}\}_{k}, such that there exists continuous subanalytic maps fk:SkQf_{k}:S_{k}\to Q with the property that for any zSkz\in S_{k} we have the equality

(3) d(z,fk(z))=d(z,Q).d(z,f_{k}(z))=d(z,Q).

In particular fk(z)=zf_{k}(z)=z for any zQz\in Q.

Moreover if SQS\setminus Q is dense in SS then there exists a subanalytic stratification of QQ by smooth manifolds such that the union of maximal strata of the stratification by the closure relation is included in fk(SkQ)\cup f_{k}(S_{k}\setminus Q). In particular fk(SkQ)\cup f_{k}(S_{k}\setminus Q) is dense in QQ.

Lemma 36.

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) be a pointed pair of closed metric subanalytic sets. Let φMDΓnb(X,p)\varphi\in MD\Gamma^{b}_{n}(X,p) be a (n,b)(n,b)-loop that has a representative {(Di,gi)}iI\{(D_{i},g_{i})\}_{i\in I} such that for every iIi\in I and for every point 𝔭:(0,1)Di\mathfrak{p}:(0,1)\to D_{i}, we have that gi𝔭g_{i}{\circ}\mathfrak{p} is bb-equivalent to a bb-point in YY. Then there is a representative {(Ej,hj)}jJ\{(E_{j},h_{j})\}_{j\in J} of the weak bb-map φ\varphi such that hj(Dj)Yh_{j}(D_{j})\subset Y for every jJj\in J.

Proof.

Apply the Lemma for S=XS=X and Q=YQ=Y. Let UU be the union of the interiors of the sets SkS_{k} predicted in Lemma 35. By subdivision and Proposition 32 φ\varphi has a representative {(Ej,hj)}jJ\{(E_{j},h^{\prime}_{j})\}_{j\in J} such that for every jJj\in J we have that hj(Ej)Sk(j)h^{\prime}_{j}(E_{j})\subset S_{k(j)} for a certain k(j)k(j). Notice that for every point 𝔭:(0,1)Ej\mathfrak{p}:(0,1)\to E_{j} the property that hj𝔭h^{\prime}_{j}{\circ}\mathfrak{p} is bb-equivalent to a point of YY is still satisfied. We define hj:=fk(j)hjh_{j}:=f_{k(j)}{\circ}h^{\prime}_{j}. Then {(Ej,hj)}jJ\{(E_{j},h^{\prime}_{j})\}_{j\in J} and {(Ej,hj)}jJ\{(E_{j},h_{j})\}_{j\in J} are bb-equivalent. ∎

Proof of Theorem 34.

The proof is an easy adaptation of the usual proof in the topological setting, taking the following caution: while in the topological proof, at some steps, a natural map InYI^{n}\to Y would be obtained, in our setting we only obtain a weak bb-map C(In)XC(I^{n})\to X that has the property of transforming any point in C(In)C(I^{n}) into a point in XX which is bb-equivalent to a point in YY. When this happens we use Lemma 36 to obtain a representative of the weak bb-map mapping C(In)C(I^{n}) into YY, and proceed with the usual steps of the topological proof.

We just give an example of this situation: in order to define the boundary homomorphism MDπkb(X,Y,𝔭)MDπk1b(Y,𝔭)MD\pi^{b}_{k}(X,Y,\mathfrak{p})\to MD\pi^{b}_{k-1}(Y,\mathfrak{p}) we start with an element φMDπkb(X,Y,𝔭)\varphi\in MD\pi^{b}_{k}(X,Y,\mathfrak{p}) and take the restriction φ|C(In1×{1}):C(In1)X\varphi|_{C(I^{n-1}\times\{1\})}:C(I^{n-1})\to X. We can not ensure that the image of φ|C(In1×{1})\varphi|_{C(I^{n-1}\times\{1\})} falls in YY in order to obtain an element of MDπk1b(Y,𝔭)MD\pi^{b}_{k-1}(Y,\mathfrak{p}) as needed, but using Lemma 36 we remedy this situation. ∎

3.6. bb-connectedness, bb-path connectedness and independence of base point.

Definition 37.

Let (X,x0,d)(X,x_{0},d) be a metric subanalytic germ. It is called bb-path connected, if for any two points 𝔮0\mathfrak{q}_{0} and 𝔮1\mathfrak{q}_{1} in XX, there is a weak bb-map η:C(I)X\eta:C(I)\to X with η(0,t)=𝔮0\eta(0,t)=\mathfrak{q}_{0} and η(1,t)=𝔮1\eta(1,t)=\mathfrak{q}_{1}.

Note that in order to have η\eta connecting 𝔮0\mathfrak{q}_{0} and 𝔮1\mathfrak{q}_{1} it is enough to have a weak bb-map connecting some points 𝔮0\mathfrak{q}_{0}^{\prime} and 𝔮1\mathfrak{q}_{1}^{\prime} bb-equivalent to the 𝔮i\mathfrak{q}_{i}’s.

The concept of bb-path connectedness is related to the concept of bb-connectedness, which was defined in Section 9 in [8] as follows:

Definition 38.

Let (X,x0)(X,x_{0}) be a metric subanalytic germ. Two connected components X1X^{1} and X2X^{2} of X{x0}X\setminus\{x_{0}\} are bb-equivalent if there exist points 𝔮1\mathfrak{q}_{1} in X1X^{1} and 𝔮2\mathfrak{q}_{2} in X2X^{2} that are bb-equivalent. The equivalence classes are called the bb-connected components of XX. The \infty-connected components are the usual connected components of X{x0}X\setminus\{x_{0}\}.

We say (X,x0,d)(X,x_{0},d) is bb-connected if it has only one bb-connected component.

They are in fact equivalent concepts:

Lemma 39.

A metric subanalytic germ (X,x0,d)(X,x_{0},d) is bb-path connected if and only if it is bb-connected.

Proof.

Assume (X,x0,d)(X,x_{0},d) is bb-connected. Let 𝔭\mathfrak{p} and 𝔭\mathfrak{p}^{\prime} be two points in XX.

Let X1X^{1}, X2X^{2}, … , XNX^{N} be the connected components of X{x0}X\setminus\{x_{0}\} (it is a finite number because X{0}X\setminus\{0\} is a subanalytic set). Moreover they are subanalytic path connected (see for example Section 3.2 in [5]). Assume 𝔭1:=𝔭\mathfrak{p}_{1}:=\mathfrak{p} belongs to X1X^{1} and 𝔭m:=𝔭\mathfrak{p}^{\prime}_{m}:=\mathfrak{p}^{\prime} to XmX^{m}. Then, by bb-connectedness, possibly after renaming the connected components XiX^{i}, there exist points 𝔭1\mathfrak{p}^{\prime}_{1}, 𝔭2\mathfrak{p}_{2}, 𝔭2\mathfrak{p}_{2}^{\prime}, …, 𝔭m1\mathfrak{p}_{m-1}, 𝔭m1\mathfrak{p}_{m-1}^{\prime}, 𝔭m\mathfrak{p}_{m} such that 𝔭k\mathfrak{p}_{k} and 𝔭k\mathfrak{p}_{k}^{\prime} belong to XkX^{k} and 𝔭kb𝔭k+1\mathfrak{p}_{k}^{\prime}\sim_{b}\mathfrak{p}_{k+1}. Let’s find weak bb-maps ηk:C(I)X\eta_{k}:C(I)\to X connecting 𝔭k\mathfrak{p}_{k} with 𝔭k\mathfrak{p}^{\prime}_{k} and ηk:C(I)X\eta^{\prime}_{k}:C(I)\to X connecting 𝔭k\mathfrak{p}_{k}^{\prime} to 𝔭k+1\mathfrak{p}_{k+1}. Then, concatenating them we would have a weak bb-map η:C(I)X\eta:C(I)\to X connecting 𝔭=𝔭1\mathfrak{p}=\mathfrak{p}_{1} and 𝔭=𝔭m\mathfrak{p}^{\prime}=\mathfrak{p}^{\prime}_{m}.

In general, given two bb-equivalent points 𝔭\mathfrak{p} and 𝔮\mathfrak{q} in XX, there always exists a weak bb-map μ:C(I)X\mu:C(I)\to X connecting them. Consider μ:={(C(0),𝔭),(C(I),f)}\mu:=\{(C(0),\mathfrak{p}),(C(I),f)\} where the map f:C(I)Xf:C(I)\to X is defined as f(y,t)=𝔮(t)f(y,t)=\mathfrak{q}(t). So in particular we can easily find the ηk\eta^{\prime}_{k} claimed in the paragraph before.

Consider now a point 𝔭:(0,ϵ)Xk\mathfrak{p}:(0,\epsilon)\to X^{k}. By Remark 9 we can choose a conical structure, that is a subanalytic homeomorphism h:XC(LX)h:X\to C(L_{X}), such that if its components are h=(h1,h2)h=(h_{1},h_{2}), where h1:XLXh_{1}:X\to L_{X} is the link component and h2:X(0,ϵ)h_{2}:X\to(0,\epsilon), then h2(x)=xh_{2}(x)=||x||. Define θ:(0,ϵ)×[0,1]LX\theta:(0,\epsilon)\times[0,1]\to L_{X} by the formulae θ(t,s):=h1(𝔭(ϵ))\theta(t,s):=h_{1}(\mathfrak{p}(\epsilon)) if t(1s)ϵt\geq(1-s)\epsilon and θ(t,s):=h1(𝔭(t+sϵ))\theta(t,s):=h_{1}(\mathfrak{p}(t+s\epsilon)) if t(1s)ϵt\leq(1-s)\epsilon. Then η:(0,ϵ)×[0,1]X\eta:(0,\epsilon)\times[0,1]\to X defined by η:=h1(θ,𝔭)\eta:=h^{-1}(\theta,||\mathfrak{p}||) is a weak bb-map connecting 𝔭\mathfrak{p} to a point 𝔮:(0,ϵ)Xk\mathfrak{q}:(0,\epsilon)\to X^{k} whose image is a retraction line of the conical structure.

Let 𝔭k\mathfrak{p}_{k} and 𝔭k\mathfrak{p}^{\prime}_{k} be points in XkX^{k}. Connect each of them as in the previous paragraph with points 𝔮k\mathfrak{q}_{k} and 𝔮k\mathfrak{q}^{\prime}_{k} whose image are retraction lines. Now using that LXkL_{X^{k}} is also subanalytic path-connected (see Section 3.2 in [5]), we construct easily a weak bb-map ηk\eta_{k} connecting 𝔮k\mathfrak{q}_{k} with 𝔮k\mathfrak{q}^{\prime}_{k}. ∎

After this, as in the topological setting we have the usual result:

Proposition 40 (Independence of base point).

Let (X,x0,d)(X,x_{0},d) be a bb-connected metric subanalytic germ. Let 𝔭1\mathfrak{p}_{1} and 𝔭2\mathfrak{p}_{2} be points in XX and let γ\gamma be a weak bb-map from C(I)C(I) to XX connecting 𝔭1\mathfrak{p}_{1} and 𝔭2\mathfrak{p}_{2}. Let γ^\hat{\gamma} be the weak bb-map from C(In)C(I^{n}) to XX defined by the formula γ^:=γρ\hat{\gamma}:=\gamma{\circ}\rho, where ρ:C(In)C(I)\rho:C(I^{n})\to C(I) is the projection ρ(y1..n,t):=(y1,t)\rho(y_{1..n},t):=(y_{1},t). Then the homomorphism

ζ:MDπnb(X,𝔭1)MDπnb(X,𝔭2)\zeta:MD\pi_{n}^{b}(X,\mathfrak{p}_{1})\rightarrow MD\pi_{n}^{b}(X,\mathfrak{p}_{2})

defined by ζ(φ):=γ^φγ^\zeta(\varphi):=\overleftarrow{\hat{\gamma}}\cdot\varphi\cdot\hat{\gamma} is an isomorphism. Moreover, its inverse is ζ1(φ):=γ^φγ^\zeta^{-1}(\varphi):=\hat{\gamma}\cdot\varphi\cdot\overleftarrow{\hat{\gamma}}.

3.7. Metric homotopy invariance

There are two notions of metric homotopy under which we have invariance. We adapt the relevant definitions from [8], and state the corresponding results. We do not include proofs because they are obvious consequences of the fact that weak bb-maps can be composed with bb-maps, or routine variations of this.

3.7.1. Metric homotopies

Definition 41 (Metric homotopy).

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) and (X,Y,x0,d,𝔭)(X^{\prime},Y^{\prime},x^{\prime}_{0},d^{\prime},\mathfrak{p}^{\prime}) be pointed pairs of metric subanalytic germs. Let

f,g:(X,Y,x0,d,𝔭)(X,Y,x0,d,𝔭)f,g:(X,Y,x_{0},d,\mathfrak{p})\to(X^{\prime},Y^{\prime},x^{\prime}_{0},d^{\prime},\mathfrak{p}^{\prime})

be Lipschitz l.v.a. subanalytic maps of pointed pairs. A continuous subanalytic map H:X×IXH:X\times I\to X^{\prime} is called a metric homotopy between ff and gg, if there is a uniform constant K0K\geq 0 such that for any ss the mapping

Hs:=H(,s):(X,Y,x0,d,𝔭)(X,Y,x0,d,𝔭)H_{s}:=H(-,s):(X,Y,x_{0},d,\mathfrak{p})\to(X^{\prime},Y^{\prime},x^{\prime}_{0},d^{\prime},\mathfrak{p}^{\prime})

is a Lipschitz l.v.a. subanalytic of pointed pairs with Lipschitz l.v.a. constant KK, and H0=fH_{0}=f and H1=gH_{1}=g. If such HH exists we say that ff and gg are metrically homotopic.

Proposition 42.

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) and (X,Y,x0,d,𝔭)(X^{\prime},Y^{\prime},x^{\prime}_{0},d^{\prime},\mathfrak{p}^{\prime}) be pointed pairs of metric subanalytic germs.

  1. (1)

    Let f,g:(X,Y,x0,d,𝔭)(X,Y,x0,d,𝔭)f,g:(X,Y,x_{0},d,\mathfrak{p})\to(X^{\prime},Y^{\prime},x^{\prime}_{0},d^{\prime},\mathfrak{p}^{\prime}) be l.v.a. subanalytic maps of pointed pairs such that there exists a continuous subanalytic mapping H:(X,Y)×I(X,Y)H:(X,Y)\times I\to(X^{\prime},Y^{\prime}) with H0=fH_{0}=f and H1=gH_{1}=g satisfying that there exists a uniform constant K>0K>0 such that for every ss, the mapping HsH_{s} is l.v.a. for the constant KK. Then we have that both fk,gk:MDπk(X,Y,𝔭)MDπk(X,Y,𝔭)f_{k}^{\infty},g_{k}^{\infty}:MD\pi^{\infty}_{k}(X,Y,\mathfrak{p})\to MD\pi^{\infty}_{k}(X^{\prime},Y^{\prime},\mathfrak{p}^{\prime}) are the same map for any kk.

  2. (2)

    Moreover, if ff and gg are metrically homotopic. Then

    fk,gk:MDπk(X,Y,𝔭)MDπk(X,Y,𝔭)f^{\star}_{k},g^{\star}_{k}:MD\pi^{\star}_{k}(X,Y,\mathfrak{p})\to MD\pi^{\star}_{k}(X^{\prime},Y^{\prime},\mathfrak{p}^{\prime})

    are equal for any kk and \star.

3.7.2. bb-homotopies

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) be a pointed pair of metric subanalytic germs. Assume that the metric dd satisfies doutdd_{out}\leq d (inner and outer metrics satisfy this). Recall from [8] (Definition 82), the definition of the metric subanalytic germ X×𝔭IX\times_{\mathfrak{p}}I. Obviously Y×𝔭IY\times_{\mathfrak{p}}I is a subgerm of X×𝔭IX\times_{\mathfrak{p}}I, and we consider in it the metric induced from X×𝔭IX\times_{\mathfrak{p}}I.

Definition 43 (bb-homotopy).

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) and (X,Y,x0,d,𝔭)(X^{\prime},Y^{\prime},x^{\prime}_{0},d^{\prime},\mathfrak{p}^{\prime}) be pointed pairs of metric subanalytic germs.metric subanalytic germs. A b-homotopy between them is a bb-map from X×𝔭IX\times_{\mathfrak{p}}I to XX^{\prime} such that there exists a representative {(Ci,fi)}iI\{(C_{i},f_{i})\}_{i\in I} satisfying

  1. (1)

    for any point 𝔭\mathfrak{p} with image inside CiY×𝔭IC_{i}\cap Y\times_{\mathfrak{p}}I, the point fi𝔭f_{i}{\circ}\mathfrak{p} is bb-equivalent to a point in YY^{\prime},

  2. (2)

    if 𝔯:(0,ϵ)X×𝔭I\mathfrak{r}:(0,\epsilon)\to X\times_{\mathfrak{p}}I projects to the point 𝔭\mathfrak{p} under the natural projection to XX, then then fi𝔯f_{i}{\circ}\mathfrak{r} is bb-equivalent to 𝔭\mathfrak{p}^{\prime}.

Proposition 44.

If there is a bb-homotopy HH with H0=fH_{0}=f and H1=gH_{1}=g, then

fkb,gkb:MDπkb(X,Y,𝔭)MDCkb(X,Y,𝔭)f_{k}^{b},g_{k}^{b}:MD\pi_{k}^{b}(X,Y,\mathfrak{p})\to MDC^{b}_{k}(X^{\prime},Y^{\prime},\mathfrak{p}^{\prime})

represent the same map for any kk.

3.8. The metric Hurewicz homomorphism

In the same way as in the topological homotopy and homology theories, for the bb-MD homology and homotopy theories there is a Hurewicz homomorphism relating them.

We recall that a simplex in (X,0)(X,0) in the Moderately Discontinuous Homology is a l.v.a. subanalytic mapping germ σ:Δ^n(X,0)\sigma:\hat{\Delta}_{n}\to(X,0) being Δ^n\hat{\Delta}_{n} the cone over the standard simplex Δn\Delta_{n}. We say that two nn-simplices σ\sigma, σ:Δ^n(X,0)\sigma^{\prime}:\hat{\Delta}_{n}\to(X,0) are bb-equivalent in the metric subanalytic germ (X,0,d)(X,0,d) if for every point pp in Δ^n\hat{\Delta}_{n} we have that σp\sigma\circ p is bb-equivalent to σp\sigma^{\prime}\circ p.

The bb-MD Chain group MDCnb(X)MDC^{b}_{n}(X) in dimension nn is the quotient of the group of formal finite sums of classes of simplices up to bb-equivalence by the homological subdivision equivalence relation (see Definition 22 in Section 3 in [8]), which essentially comes from identifying a simplex σ:Δ^n(X,0)\sigma:\hat{\Delta}_{n}\to(X,0) with any formal sum iσρi\sum_{i}\sigma\circ\rho_{i} where the ρi:Δ^nΔ^n\rho_{i}:\hat{\Delta}_{n}\to\hat{\Delta}_{n} are orientation preserving l.v.a. subanalytic parametrizations of the simplices of a finite subanalytic triangulation of Δ^n\hat{\Delta}_{n}.

With this in mind, we define a map from (n,b)(n,b)-loops, or more generally weak bb-maps, to the bb-MD chains of a pointed metric subanalytic germ ((X,x0,d),p)((X,x_{0},d),p):

(4) ζn,b:{φ:φ is a weak b-map from C(In) to X}MDCnb(X;)\zeta_{n,b}:\{\varphi:\varphi\text{ is a weak }b\text{-map from }C(I^{n})\text{ to }X\}\to MDC_{n}^{b}(X;\mathbb{Z})

as follows: given φ={(Cj,fj)}jJ\varphi=\{(C_{j},f_{j})\}_{j\in J} be a weak bb-map from C(In)C(I^{n}) to XX we define the map ζ\zeta by the formula

(5) ζn,b(φ):=kKfr(k)ρk\zeta_{n,b}(\varphi):=\sum_{k\in K}f_{r(k)}{\circ}\rho_{k}

where {ρk:Δ^n(C(In),0)}kK\{\rho_{k}:\hat{\Delta}_{n}\to(C(I^{n}),0)\}_{k\in K} is an orientation preserving homological subdivision (recall Definition 18 of [8]) of C(In)C(I^{n}) whose associated triangulation is compatible with {Cj}jJ\{C_{j}\}_{j\in J} and r(k)Jr(k)\in J is such that the image of ρk\rho_{k} is contained in Cr(k)C_{r(k)} for every kk.

Lemma 45.

Let (X,x0,d,𝔭)(X,x_{0},d,\mathfrak{p}) be a pointed metric subanalytic germ. Let ζn,b\zeta_{n,b} be as defined above. Then ζn,b\zeta_{n,b} has the following properties:

  1. (1)

    The map ζn,b\zeta_{n,b} is well-defined independent of the choice of the homological subdivision.

  2. (2)

    If φ1\varphi_{1} and φ2\varphi_{2} can be concatenated then we have ζn,b(φ1φ2)=ζn,b(φ1)+ζn,b(φ2)\zeta_{n,b}(\varphi_{1}\cdot\varphi_{2})=\zeta_{n,b}(\varphi_{1})+\zeta_{n,b}(\varphi_{2}).

  3. (3)

    We have ζn,b(φ)=ζn,b(φ)\zeta_{n,b}(\overleftarrow{\varphi})=-\zeta_{n,b}(\varphi).

Proof.

All the stated properties follow from the homological subdivision equivalence in MDCnb((X,Im(𝔭));)MDC_{n}^{b}((X,\text{Im}(\mathfrak{p}));\mathbb{Z}). In particular, for property (3) recall Remark 20 of [8]. ∎

Proposition 46 (Metric Hurewicz homomorphism).

Let ((X,x0,d,𝔭)((X,x_{0},d,\mathfrak{p}) be a pointed metric subanalytic germ. Let b(0,]b\in(0,\infty] and nn\in\mathbb{N}. Then the restriction of ζn,b\zeta_{n,b} to the space of (n,b)(n,b)-loops induces a homomorphism

ζn,b¯:MDπnb(X,𝔭)MDHnb(X;),\overline{\zeta_{n,b}}:MD\pi_{n}^{b}(X,\mathfrak{p})\to MDH_{n}^{b}(X;\mathbb{Z}),

which we call the Hurewicz morphism.

Proof.

By Lemma 45, if ζn,b¯\overline{\zeta_{n,b}} is well-defined, then it is a homomorphism.

We consider the relative homology complex MDCnb(X,Im(𝔭);)MDC_{n}^{b}(X,\text{Im}(\mathfrak{p});\mathbb{Z}) (See section 3.5 of [8]). Since MDHnb(X;)MDHnb(X,Im(𝔭);)MDH_{n}^{b}(X;\mathbb{Z})\to MDH_{n}^{b}(X,\text{Im}(\mathfrak{p});\mathbb{Z}) is an isomorphism for any n>0n>0 we can consider the composition of ζn,b\zeta_{n,b} with the natural projection to MDCnb(X,Im(𝔭);)MDC_{n}^{b}(X,\text{Im}(\mathfrak{p});\mathbb{Z}) instead of ζn,b\zeta_{n,b} itself. Then the image of any nn-loop in MDπnb(X,𝔭)MD\pi_{n}^{b}(X,\mathfrak{p}) is obviously a cycle, and homotopic nn-loops give rise to homologous cycles. Moreover given the constant bb-loop c𝔭,nc_{\mathfrak{p},n} we have ζn,b(c𝔭,n)=0\zeta_{n,b}(c_{\mathfrak{p},n})=0. ∎

Theorem 47 (Metric Hurewicz isomorphism).

Let ((X,x0,d,𝔭)((X,x_{0},d,\mathfrak{p}) be a pointed metric subanalytic germ. Let b(0,]b\in(0,\infty]. If XX is bb-connected then the group MDH1b(X;)MDH_{1}^{b}(X;{\mathbb{Z}}) is the abelianization of MDπ1b(X,𝔭)MD\pi_{1}^{b}(X,\mathfrak{p}), and the Hurewicz homomorphism is the abelianization map.

Assume that MDπ1b(X,𝔭)MD\pi_{1}^{b}(X,\mathfrak{p}) is trivial and that dd is either the inner or the outer metric. If MDπkb(X,𝔭)MD\pi_{k}^{b}(X,\mathfrak{p}) is trivial for k<nk<n then ζn,b¯:MDπnb(X,𝔭)MDHnb(X;)\overline{\zeta_{n,b}}:MD\pi_{n}^{b}(X,\mathfrak{p})\to MDH_{n}^{b}(X;\mathbb{Z}) is a group isomorphism.

Proof.

The proof of the fundamental group part is an adaptation of the usual topological proof (see [9]) to our setting, and it involves no new ideas. The reader interested in the details may consult [10]. The higher homotopy groups case is harder and is provided at the end of Section 5.2. The proof in Section 5.2 also contains the fundamental group statement for the inner and the outer metrics.

4. The metric Seifert van Kampen Theorem for the MD-Fundamental groupoid.

4.1. The MD Fundamental Groupoid.

Along this section we use the following terminology to study a metric subanalytic germ (X,0,d)(X,0,d). Given two bb-points 𝔭\mathfrak{p} and 𝔮\mathfrak{q} in XX, a bb-path from 𝔭\mathfrak{p} to 𝔮\mathfrak{q} is a weak bb-maps γ:C(I)(X,0)\gamma:C(I)\to(X,0) such that γ(0,_)b𝔭\gamma(0,\_)\sim_{b}\mathfrak{p} and γ(1,_)b𝔮\gamma(1,\_)\sim_{b}\mathfrak{q}. Two bb-paths γ\gamma and β\beta from 𝔭\mathfrak{p} to 𝔮\mathfrak{q} are bb-homotopy equivalent if there exists a homotopy η:C(I×I)X\eta:C(I\times I)\to X relative to the extremes C({0,1}×I)C(\{0,1\}\times I) such that η|C(I×{0})\eta|_{C(I\times\{0\})} is bb-equivalent to γ\gamma and η|C(I×{1})}\eta|_{C(I\times\{1\})\}} is bb-equivalent to β\beta.

Definition 48.

Let (X,0,d)(X,0,d) be a metric subanalytic germ. The bb-MD Fundamental Groupoid of XX is the category, denoted by MDπ1b(X,d)MD\pi_{1}^{b}(X,d) or MDπ1b(X)MD\pi^{b}_{1}(X), whose objects are the bb-points of (X,0)(X,0) and whose morphisms from 𝔭\mathfrak{p} to 𝔮\mathfrak{q} are the bb-homotopy classes of bb-paths from 𝔭\mathfrak{p} to 𝔮\mathfrak{q} relative to the extremes.

Let 𝔓\mathfrak{P} be a subset of bb-points in XX. We denote by MDπ1b(X,d,𝔓)MD\pi_{1}^{b}(X,d,\mathfrak{P}) or MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) the full subcategory of MDπ1b(X)MD\pi_{1}^{b}(X) whose set of objects is 𝔓\mathfrak{P}.

Composition in MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) is given by concatenation (see Definition 23) and every morphism is invertible: the inverse of the bb-path class of γ\gamma has γ\overleftarrow{\gamma} as a representative (see Notation 25). When 𝔓\mathfrak{P} is a single point {𝔭}\{\mathfrak{p}\} we recover the bb-MD fundamental group (see Proposition 26).

Analogously to Proposition 28 we have the following:

Proposition 49.

For every b(0,+)b\in(0,+\infty), the bb-MD Fundamental Grupoid MDπ1bMD\pi^{b}_{1} is a functor from the category of metric subanalytic germs (with both Lipschitz l.v.a. maps as morphism or bb-maps) to the category of groupoids. Moreover, for bbb\geq b^{\prime} there are groupoid morphisms, that we denote as in (2)

(6) ηb,b:MDπ1b(X,𝔓)MDπ1b(X,𝔓).\eta_{b,b^{\prime}}:MD\pi^{b}_{1}(X,\mathfrak{P})\to MD\pi^{b^{\prime}}_{1}(X,\mathfrak{P}).

So, we have the bb-MD Fundamental Groupoid MDπ1bMD\pi^{b}_{1} as a functor from the category of metric subanalytic germs (with both Lipschitz l.v.a. maps as morphism or bb-maps) to the category 𝔹𝒢𝒫{\mathbb{B}}-{\mathcal{G}}{\mathcal{P}}.

We state a basic observation that will be used later:

Lemma 50.

Let (X,0,d)(X,0,d) be a metric subanalytic germ. Let 𝔓\mathfrak{P} be a set of bb-points of XX. Let γ:C(I)X\gamma:C(I)\to X be a bb-path with extremes in points of 𝔓\mathfrak{P}. Then, given any subanalytic homeomorphism h:C(I)C(I)h:C(I)\to C(I) which resticts to the identity in C({0})C(\{0\}) and C({1}C(\{1\} we have that [γh]=[γ][\gamma\circ h]=[\gamma] in MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}).

Proof.

Consider the homotopy H:C(I)×IC(I)H:C(I)\times I\to C(I) defined by H(yt,t,s):=(1s)(yt,t)+sh(yt,t)H(yt,t,s):=(1-s)(yt,t)+sh(yt,t) and compose it with γ\gamma. ∎

4.2. The metric Seifert- van Kampen Theorem.

In this section we consider a covering {Ui}i\{U_{i}\}_{i} of a metric subanalytic germ (X,0,d)(X,0,d). We will always considered the subsets UiU_{i} with metrics did_{i} where did_{i} are defined by one of the following cases:

  • (a)

    either the did_{i}’s are the restriction metrics d|Uid|_{U_{i}}

  • (b)

    or the did_{i}’s are the inner metrics in UiU_{i} induced by the infimum of lengths of rectifiable paths in (X,d)(X,d). A particular case is when dd is the outer or inner metric in XX induced by the euclidean metric in n{\mathbb{R}}^{n} and then the did_{i}’s are the inner metrics induced in every UiU_{i}.

Definition 51.

Let 𝒰={Ui}iI{\mathcal{U}}=\{U_{i}\}_{i\in I} be a finite cover of a metric subanalytic germ (X,0,d)(X,0,d) by subanalytic open subsets UiU_{i} endowed with metrics did_{i} (all either of type (a) o (b) above). Let 𝔓\mathfrak{P} be a set of bb-points in XX. The pair (𝒰,𝔓)({\mathcal{U}},\mathfrak{P}) satisfies condition

  • ()b(*)_{b}

    if for every bb-equivalent points 𝔮i,𝔮j\mathfrak{q}_{i},\mathfrak{q}_{j} and 𝔮k\mathfrak{q}_{k} in UiU_{i}, UjU_{j} and UkU_{k} respectively (where ii, jj and kk may coincide), there exist

    • a point 𝔭\mathfrak{p} in 𝔓\mathfrak{P},

    • a bb-path δv:C(I)(Uv,dv)\delta_{v}:C(I)\to(U_{v},d_{v}) joining 𝔮v\mathfrak{q}_{v} with 𝔭\mathfrak{p} for every v=i,j,kv=i,j,k ,

    • weak bb-homotopies μvw:C(I2)(Uv,dv)\mu_{vw}:C(I^{2})\to(U_{v},d_{v}) for every v,w{i,j,k}v,\ w\in\{i,j,k\} relative to the extremes 𝔭\mathfrak{p} and 𝔮v\mathfrak{q}_{v}, with μvw|C(I×{0})bδv\mu_{vw}|_{C(I\times\{0\})}\sim_{b}\delta_{v} in (Uv,dv)(U_{v},d_{v}) and such that for every v,w=i,j,kv,w=i,j,k the bb-paths μvw|C(I×{1})\mu_{vw}|_{C(I\times\{1\})} and μwv|C(I×{1})\mu_{wv}|_{C(I\times\{1\})} are bb-equivalent as bb-maps in (X,d)(X,d).

Notation 52.

Given a metric subanalytic germ (X,0,d)(X,0,d), a subanalytic subgerm UXU\subset X and a set 𝔓\mathfrak{P} of bb-points in XX we denote by 𝔓U\mathfrak{P}\cap U the subset of points in 𝔓\mathfrak{P} which are bb-equivalent in (X,d)(X,d) to a point in UU. Then we can consider π1b(U,𝔓U)\pi_{1}^{b}(U,\mathfrak{P}\cap U) which we will denote simply by π1b(U,𝔓)\pi_{1}^{b}(U,\mathfrak{P}). Given VUXV\subset U\subset X considered with metrics dVd_{V} and dUd_{U} of the same type with respect to Definition 51, functoriality gives rise to a morphism of groupoids

ιV,U:MDπ1b(V,𝔓)π1b(U,𝔓).\iota_{V,U}:MD\pi_{1}^{b}(V,\mathfrak{P})\to\pi_{1}^{b}(U,\mathfrak{P}).
Definition 53.

Let (X,0,d)(X,0,d) be a metric subanalytic germ, 𝒰={Ui}iI{\mathcal{U}}=\{U_{i}\}_{i\in I} a finite cover and 𝔓\mathfrak{P} a set of bb-points in XX. A groupoid KK and a collection of groupoid morphisms ai:MDπ1b(Ui,𝔓)Ka_{i}:MD\pi_{1}^{b}(U_{i},\mathfrak{P})\to K with iIi\in I satisfy property

  • ()b(\dagger)_{b}

    if for any [γi]MorMDπ1b(Ui,𝔓)(𝔭,𝔮)[\gamma_{i}]\in Mor_{MD\pi^{b}_{1}(U_{i},\mathfrak{P})}(\mathfrak{p},\mathfrak{q}) and [γj]MorMDπ1b(Uj,𝔓)(𝔭,𝔮)[\gamma_{j}]\in Mor_{MD\pi^{b}_{1}(U_{j},\mathfrak{P})}(\mathfrak{p},\mathfrak{q}) having representatives γi\gamma_{i} en UiU_{i} and γj\gamma_{j} en UjU_{j} which are equivalent as bb-maps in XX we have the equality ai([γi])=aj([γj)]a_{i}([\gamma_{i}])=a_{j}([\gamma_{j})].

Theorem 54 (A universal property characterizing the MD-fundamental groupoid).

Let (X,0,d)(X,0,d) be a metric subanalytic germ and (𝒰,𝔓)({\mathcal{U}},\mathfrak{P}) be a finite cover and a set of bb-points in XX satisfying condition ()(*). There is a unique (up to groupoid isomorphism) groupoid LL and groupoid morphisms bi:MDπ1(Ui,𝔓)Lb_{i}:MD\pi_{1}(U_{i},\mathfrak{P})\to L satisfying property ()b(\dagger)_{b} such that for any other groupoid KK and groupoid morphisms κi:MDπ1b(Ui,𝔓)K\kappa_{i}:MD\pi_{1}^{b}(U_{i},\mathfrak{P})\to K with the property ()b(\dagger)_{b} there exists a unique groupoid morphism κ:LK\kappa:L\to K such that κbi=κi\kappa{\circ}b_{i}=\kappa_{i} for any iIi\in I. Moreover MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) and the morphisms ιUi,X\iota_{U_{i},X} coincide with LL and bib_{i} for any iIi\in I.

In other words and informally: the fundamental groupoid MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) is the unique initial object of the category of groupoids having property ()b(\dagger)_{b}.

Proof.

Given the existence, the uniqueness is obvious. It is also clear that MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) and ιUi,X\iota_{U_{i},X} satisfy property ()(\dagger). Consider a groupoid KK and groupoid morphisms κi:MDπ1b(Ui,𝔓)K\kappa_{i}:MD\pi_{1}^{b}(U_{i},\mathfrak{P})\to K with property ()(\dagger). We have to find the predicted groupoid morphism κ:MDπ1b(X,𝔓)K\kappa:MD\pi_{1}^{b}(X,\mathfrak{P})\to K and to prove that κ\kappa is unique.

The definition and uniqueness of κ\kappa at the level of objects follows from condition ()(\dagger) applied to the constant loops (that are the unit morphisms in the fundamental groupoid), the bijection between objects and unit elements of the isotropy groups of a groupoid and the fact that groupoid morphisms preserve unit elements.

In the rest of the proof we prove that there is a unique possible definition of κ\kappa at the level of morphisms. We adapt the line of the proof of Seifert-van Kampen for groupoids in [4] (see also [9] for the fundamental group case) in the topological setting, with some additional non-trivial arguments.

Given 𝔭,𝔮𝔓\mathfrak{p},\mathfrak{q}\in\mathfrak{P} we define the map κ:MorMDπ1b(X,𝔓)(𝔭,𝔮)MorK(κ(𝔭),κ(𝔮))\kappa:Mor_{MD\pi_{1}^{b}(X,\mathfrak{P})}(\mathfrak{p},\mathfrak{q})\to Mor_{K}(\kappa(\mathfrak{p}),\kappa(\mathfrak{q})). Given an element [γ]MorMDπ1b(X,𝔓)(𝔭,𝔮)[\gamma]\in Mor_{MD\pi_{1}^{b}(X,\mathfrak{P})}(\mathfrak{p},\mathfrak{q}), if

(7) [γ]=[γ1][γr][\gamma]=[\gamma_{1}]\centerdot...\centerdot[\gamma_{r}]

where[γl][\gamma_{l}] are elements of MorMDπ1b(Uil,𝔓)(𝔭l,𝔭l+1)Mor_{MD\pi_{1}^{b}(U_{i_{l}},\mathfrak{P})}(\mathfrak{p}_{l},\mathfrak{p}_{l+1}) for certain ilIi_{l}\in I and 𝔭l\mathfrak{p}_{l}, 𝔭l+1𝔓\mathfrak{p}_{l+1}\in\mathfrak{P}, then necessarily

(8) κ([γ]):=κi1([γ1])κir([γr]).\kappa([\gamma]):=\kappa_{i_{1}}([\gamma_{1}])\centerdot...\centerdot\kappa_{i_{r}}([\gamma_{r}]).

Notice that if any [γl][\gamma_{l}] has a bb-map representative γl\gamma_{l}^{\prime} with image contained in a different UilU_{i_{l}^{\prime}}, then property ()(\dagger) of the system of morphisms {κi}iI\{\kappa_{i}\}_{i\in I} shows the equality κil([γl])=κil([γl])\kappa_{i_{l}}([\gamma_{l}])=\kappa_{i_{l}^{\prime}}([\gamma_{l}^{\prime}]) so the value obtained for κ([γ])\kappa([\gamma]) is independent of the choice of the subset UilU_{i_{l}}.

Notice that any bb-loop [γ][\gamma] admits an expression as (7). Indeed, let {(Cj,fj)}jJ\{(C_{j},f_{j})\}_{j\in J} be a representative of γ\gamma such that such that for each jj there is a iji_{j} such that fj(Cj)Uijf_{j}(C_{j})\subset U_{i_{j}} (this can be achieved by taking the refinement {Cjfj1(Ui)}i,j\{C_{j}\cap f_{j}^{-1}(U_{i})\}_{i,j} of any covering {Cj}j\{C_{j}\}_{j}). By the existence of subanalytic triangulations and Remark 9 there is a triangulation h:|K|C(I)h:|K|\to C(I) compatible with the decomposition by {Cj}jJ\{C_{j}\}_{j}\in J and preserving tt-levels CiC_{i}. Applying Lemma 50 we have that γh\gamma\circ h is a representative of the homotopy class [γ][\gamma], and it is clear that it is a concatenation of bb-paths γ0γr\gamma_{0}\cdot...\gamma_{r}, each of which has image contained in a subset UiU_{i} of the cover. Now we modify each γj\gamma_{j} so that its starting and ending points belong to 𝔓\mathfrak{P}. Since the ending point 𝔢j\mathfrak{e}_{j} of γj\gamma_{j} and the starting point 𝔰j+1\mathfrak{s}_{j+1} of γj+1\gamma_{j+1} coincide as bb-points in XX, by condition ()(*) there exists a point 𝔭j+1𝔓\mathfrak{p}_{j+1}\in\mathfrak{P}, bb-paths δj+10\delta_{j+1}^{0} and δj+11\delta_{j+1}^{1} connecting 𝔢j\mathfrak{e}_{j} with 𝔭j+1\mathfrak{p}_{j+1} inside UijU_{i_{j}} and 𝔰j+1\mathfrak{s}_{j+1} with 𝔭j+1\mathfrak{p}_{j+1} inside Uij+1U_{i_{j+1}} respectively, and so that δj+10δj+11\delta_{j+1}^{0}\cdot\overleftarrow{\delta_{j+1}^{1}} is weak bb-homotopic to a constant bb-path. Then we have

[γ]=[γ0δ10][δ11γ1δ20][δr1γr],[\gamma]=[\gamma_{0}\cdot\delta_{1}^{0}]\cdot[\overleftarrow{\delta_{1}^{1}}\cdot\gamma_{1}\cdot\delta_{2}^{0}]\cdot...\cdot[\overleftarrow{\delta_{r}^{1}}\cdot\gamma_{r}],

which is the needed expression of the form (7).

In order to finish the proof, we have to see that for any two expressions of an element [γ][\gamma] as in (7)(\ref{eq:concat}), the values κ([γ])\kappa([\gamma]) induced as in (8) are the same: as a conclusion, we get that κ\kappa is well defined.

Consider two expressions [γ]=[γ1γr]=[γ1γr][\gamma]=[\gamma_{1}\cdot...\cdot\gamma_{r}]=[\gamma^{\prime}_{1}\cdot...\cdot\gamma^{\prime}_{r^{\prime}}] as in (7)(\ref{eq:concat}). Let η\eta be a bb-homotopy connecting γ1γr\gamma_{1}\cdot...\cdot\gamma_{r} and γ1γr\gamma^{\prime}_{1}\cdot...\cdot\gamma^{\prime}_{r^{\prime}} relative to the extremes, given by a representative {(Dk,gk)}kA\{(D_{k},g_{k})\}_{k\in A} with the DkDlD_{k}\cap D_{l} with empty interior and such that gl(Dl)Ui(l)g_{l}(D_{l})\subset U_{i(l)} for certain i(l)Ii(l)\in I. Since we have the bb-equivalences η|C(I×{0})bγ1γr\eta|_{C(I\times\{0\})}\sim_{b}\gamma_{1}\cdot...\cdot\gamma_{r} and η|C(I×{1})bγ1γr,\eta|_{C(I\times\{1\})}\sim_{b}\gamma^{\prime}_{1}\cdot...\cdot\gamma^{\prime}_{r^{\prime}}, after a refinement of the partition, we may assume that {(Dk,gk)}kA\{(D_{k},g_{k})\}_{k\in A} refines the decompositions induced in C(I×{0})C(I\times\{0\}) and C(I×{1})C(I\times\{1\}) by the concatenation expressions. Next we will modify the homotopy η\eta and the decomposition {(Dk,gk)}kA\{(D_{k},g_{k})\}_{k\in A} in several steps so that

  1. (i)

    the DkD_{k} are cones over convex polygons in I2I_{2} that only intersect along faces,

  2. (ii)

    no more than three DkD_{k} meet in a point.

The condition (ii) is crucial because we will need to apply property (*), which is formulated for coverings allowing only 33-fold intersections. The first condition is very convenient to easily express certain restrictions of the homotopy (over the faces of the DkD_{k}’s) as a concatenation of paths.

We consider a triangulation τ:C(|K|)C(I2)\tau:C(|K|)\to C(I^{2}) where KK is a simplicial complex (and |K|=I2|K|=I^{2}), adapted to the natural strata of C(I2)C(I^{2}) and to the {Dk}k\{D_{k}\}_{k} and that preserves tt-levels. Then, the homotopy η~:=ητ\widetilde{\eta}:=\eta\circ\tau is a weak bb-homotopy between the concatenations of reparametrizations γ~i\widetilde{\gamma}_{i} of the bb-paths γi\gamma_{i} and γ~i\widetilde{\gamma}^{\prime}_{i} of γi\gamma^{\prime}_{i}. Let Uj(i)U_{j(i)} be a subset of the cover containing the image of γi\gamma_{i}. By Lemma 50 the path [γi][\gamma_{i}] represents the same element in MDπ1b(Uj(i),𝔓)MD\pi_{1}^{b}(U_{j(i)},\mathfrak{P}) than γ~i\widetilde{\gamma}_{i}. So, by ()(\dagger) we have

(9) κ[γ1γ2γr]=κ[γ~1γ~r]andκ[γ1γr]=κ[γ~1γ~r].\kappa[\gamma_{1}\cdot\gamma_{2}\cdot...\cdot\gamma_{r}]=\kappa[\widetilde{\gamma}_{1}\cdot...\cdot\widetilde{\gamma}_{r}]\ \ \text{and}\ \ \kappa[\gamma^{\prime}_{1}\cdot...\cdot\gamma^{\prime}_{r^{\prime}}]=\kappa[\widetilde{\gamma}^{\prime}_{1}\cdot...\cdot\widetilde{\gamma}^{\prime}_{r^{\prime}}].

So, we have to check the equality of the two second terms of each of the previous equalities, and we can work with the homotopy η~\widetilde{\eta}. We denote by {Tl}l\{T_{l}\}_{l} the family of 2-dimensional simplices of KK, which decompose |K|=I2|K|=I^{2}.

In case there are vertices in KK that meet more than three 2-dimensional simplices, we do the following modifications of the partition {Tl}l\{T_{l}\}_{l} of I2I^{2} and of η~\widetilde{\eta}. For every vertex vv of valency m>3m>3 we consider a convex mm-gonal piece GvG_{v} (the union of mm small triangular pieces each of them inside one of the adjacent 2-dimensional simplices) as in Figure 2. We choose them small enough so that they are disjoint as in the figure. We consider the partition 𝒫{\mathcal{P}} of I2I^{2} given by {Gv}val(v)>3\{G_{v}\}_{val(v)>3} and {T~i}i\{\widetilde{T}_{i}\}_{i} with T~i:=TivGv\widetilde{T}_{i}:=T_{i}\setminus\cup_{v}G_{v}. We consider the conical partition C(𝒫)C({\mathcal{P}}) induced on C(I2)C(I^{2}). We now modify the homotopy η~\widetilde{\eta}. Consider the continuous mapping τ2:𝒫|K|\tau_{2}:{\mathcal{P}}\to|K| as follows:

  • the restriction to every GvG_{v} collapses, to vv

  • it is the identity in the rest of the vertices,

  • the restriction to every Ti~\widetilde{T_{i}} has image TiT_{i} and is any subanalytic map satisfying the following conditions: it is a homeomorphism of the interiors and at the boundary it interpolates affinely the definition provided in previous 2 items.

Refer to caption
Figure 2. The refinement 𝒫{\mathcal{P}} of the partition |K||K| of I2I^{2}.

We also denote by τ2\tau_{2} the induced conical map from C(𝒫)C({\mathcal{P}}) to C(|K|)C(|K|) given by (y,t)(τ2(y),t)(y,t)\mapsto(\tau_{2}(y),t). We define a bb-homotopy ξ:=η~τ2\xi:=\widetilde{\eta}\circ\tau_{2} as follows. For every region C(T~i)C(\widetilde{T}_{i}) there is a DkD_{k} such that τ(C(Ti))Dk\tau(C(T_{i}))\subset D_{k}. Then we define ξ|C(T~i}):=gkττ2\xi|_{C(\widetilde{T}_{i}\})}:=g_{k}{\circ}\tau{\circ}\tau_{2}. At a piece C(Gv)C(G_{v}) there are several DkD_{k}’s such that τ2(C(Gv))Dk\tau_{2}(C(G_{v}))\subset D_{k}. Choose one of them and define ξ:=gkττ2\xi:=g_{k}{\circ}\tau{\circ}\tau_{2} there. The reader may check easily that we obtain a bb-map by this procedure. By abuse of notation we denote also by {Dk}k\{D_{k}\}_{k} its associated covering {C(T~i)}i{C(Gv)}\{C(\widetilde{T}_{i})\}_{i}\cup\{C(G_{v})\}.

The replacement of the homotopy η~\widetilde{\eta} by ξ\xi performed above induces a replacement of the concatenations [γ~1γ~r][\widetilde{\gamma}_{1}\cdot...\cdot\widetilde{\gamma}_{r}] and [γ~1γ~r][\widetilde{\gamma}^{\prime}_{1}\cdot...\cdot\widetilde{\gamma}^{\prime}_{r^{\prime}}] to which η~\widetilde{\eta} resticts to C(I×{0})C(I\times\{0\}) and C(I×{1})C(I\times\{1\}). The restriction of ξ\xi to C(I×{0})C(I\times\{0\}) is the concatenation [γ~1γ~r][\widetilde{\gamma}_{1}\cdot...\cdot\widetilde{\gamma}_{r}] replaced by the same concatenation in which several constant bb-loops (corresponding to the GvG_{v} regions meeting C(I×{0})C(I\times\{0\})) are inserted. Similarly for [γ~1γ~r][\widetilde{\gamma}^{\prime}_{1}\cdot...\cdot\widetilde{\gamma}^{\prime}_{r^{\prime}}]. Since the value of κ\kappa is unchanged under this operation we can safely assume that the homotopy connecting the concatenations is ξ\xi, which satisfies (i) and (ii).

To finish the proof we need to check the equality

(10) κ[γ~1γ~r]=κ[γ~1γ~r].\kappa[\widetilde{\gamma}_{1}\cdot...\cdot\widetilde{\gamma}_{r}]=\kappa[\widetilde{\gamma}^{\prime}_{1}\cdot...\cdot\widetilde{\gamma}^{\prime}_{r^{\prime}}].

Now, we modify ξ\xi, analogously to the classical proof [4], so that it sends all the meeting points of 3 cells in the decomposition C(𝒫)C({\mathcal{P}}) to points in 𝔓\mathfrak{P}. Note that in the case of points over C(I×{0})C(I\times\{0\}) or C(I×{1})C(I\times\{1\}) they are already points in 𝔓\mathfrak{P}, so we do not need any modification there. Let 𝔮\mathfrak{q} be the point in C(I2)C(I^{2}) whose image is the intersection of 33 cells Du,Dv,DwD_{u},D_{v},D_{w} of the decomposition C(𝒫)C({\mathcal{P}}). Recall that the points gu𝔮g_{u}{\circ}\mathfrak{q}, gv𝔮g_{v}{\circ}\mathfrak{q} and gw𝔮g_{w}{\circ}\mathfrak{q} are pairwise bb-equivalent. By condition (*) there exist

  • a bb-point 𝔭𝔓\mathfrak{p}\in\mathfrak{P},

  • bb-paths δu:C(I)Uj(u)\delta_{u}:C(I)\to U_{j(u)} joining gu𝔮g_{u}{\circ}\mathfrak{q} with 𝔭\mathfrak{p}, δv:C(I)Uj(v)\delta_{v}:C(I)\to U_{j(v)} joining gv𝔮g_{v}{\circ}\mathfrak{q} with 𝔭\mathfrak{p} and δw:C(I)Uj(w)\delta_{w}:C(I)\to U_{j(w)} joining gw𝔮g_{w}{\circ}\mathfrak{q} with 𝔭\mathfrak{p},

  • weak bb-homotopies μvw:C(I2)Uj(v)\mu_{vw}:C(I^{2})\to U_{j(v)} for every pair v,w{u,v,w}2v,w\in\{u,v,w\}^{2} of different indexes with μvw|C(I×{0})bδv\mu_{vw}|_{C(I\times\{0\})}\sim_{b}\delta_{v} and such that the condition in (*) holds.

We modify the bb-homotopy ξ\xi in a neighbourhood of every intersection of every (two or) three cells DkD_{k} according to Figure 3, gluing the bb-homotopies μij\mu_{ij} and some composition of ξ\xi with some collapsing mapping similar to the τ2\tau_{2} used above. This is just as the standard procedure in classical topology, plus the observation that condition (*) is what we need so that we actually obtain a weak bb-map. We leave the details to the reader. We call ξ\xi^{\prime} the resulting bb-homotopy.

Refer to caption
Figure 3. Modification around the vertices of the decomposition C(𝒫)C({\mathcal{P}}), using hypothesis (*), to obtain the weak bb-homotopy ξ\xi^{\prime}.
Refer to caption
Figure 4. Possible choice for the polygonal line B4B_{4} for the example in the previous figures.

Now we adapt the procedure of the classical proof (see [9]). We choose a level t=t0t=t_{0} in C(I2)C(I^{2}) and denote by Dj0D_{j}^{0} the intersection of DjD_{j} with that level. Considering a numbering D10,.,Dm0D^{0}_{1},....,D^{0}_{m} of the polygons which cover I2I^{2} such that for any nmn\leq m the union Qn:=(I×{1})i=1nDn0Q_{n}:=(I\times\{1\})\cup\cup_{i=1}^{n}D^{0}_{n} is contractible. For any nn we define BnB_{n} the lower boundary of QnQ_{n}, that is the connected polygonal line containing the points (0,1)(0,1) and (1,1)(1,1) defined as the closure of the boundary of [0,1]×(Bn[0,1]×(1,+))[0,1]\times{\mathbb{R}}\setminus(B_{n}\cup[0,1]\times(1,+\infty)): see Figure 4.

Then we have concatenation expressions

ξ|Bn=α1αlαr,\xi^{\prime}|_{B_{n}}=\alpha_{1}\cdot...\alpha_{l}\cdot...\cdot\alpha_{r},
ξ|Bn+1=α1αlαr,\xi^{\prime}|_{B_{n+1}}=\alpha_{1}\cdot...\cdot\alpha^{\prime}_{l}\cdot...\cdot\alpha_{r},

where each αi\alpha_{i} is a b-path that connects two bb-points of 𝔓\mathfrak{P}, bounds partially some DkD_{k}, has image in a subset UiU_{i} and αl\alpha_{l} and αl\alpha^{\prime}_{l} are the bb-paths whose union is the boundary of the region Dn+1D_{n+1} (that is C(Dn+10C(D_{n+1}^{0}). Then gi(Dn+1)g_{i}(D_{n+1}) is contained in a subset UjU_{j}. Hence [αl][\alpha_{l}] and [αl][\alpha^{\prime}_{l}] are equal in MDπ1b(Uj,𝔓)MD\pi_{1}^{b}(U_{j},\mathfrak{P}). This shows that for every nn

(11) κ([ξ|Bn])=κ([ξ|Bn+1]).\kappa([\xi^{\prime}|_{B_{n}}])=\kappa([\xi^{\prime}|_{B_{n+1}}]).

Then, we conclude (10) since it is clear that κ([ξ|B0])=κ([ξ|I×{1}])\kappa([\xi^{\prime}|_{B_{0}}])=\kappa([\xi|_{I\times\{1\}}]) and κ([ξ|Bm])=κ([ξ|I×{0}])\kappa([\xi^{\prime}|_{B_{m}}])=\kappa([\xi|_{I\times\{0\}}]) and they are equal by (11).

Let (X,0,d)(X,0,d) be a metric subanalytic germ and (𝒰,𝔓)({\mathcal{U}},\mathfrak{P}) be a finite cover and a set of bb-points in XX satisfying condition ()(*). We wish to compare MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) with the colimit colim{MDπ1b(Ui1,,ir,𝔓)}i1.ir\mathrm{colim}\{MD\pi_{1}^{b}(U_{i_{1},...,i_{r}},\mathfrak{P})\}_{i_{1}....i_{r}}. By the universal property that we have just proved for MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) and the one defining the colimit we obtain a natural morphism of groupoids

(12) σ:colim{MDπ1b(Ui1,,ir,𝔓)}i1.irMDπ1b(X,𝔓).\sigma:\mathrm{colim}\{MD\pi_{1}^{b}(U_{i_{1},...,i_{r}},\mathfrak{P})\}_{i_{1}....i_{r}}\to MD\pi_{1}^{b}(X,\mathfrak{P}).

We wish to find conditions ensuring that σ\sigma is an isomorphism.

Theorem 55.

Let (X,0,d)(X,0,d) be a metric subanalytic germ and (𝒰,𝔓)({\mathcal{U}},\mathfrak{P}) be a finite cover and a set of bb-points in XX satisfying condition ()b(*)_{b}. Then σ\sigma is surjective at the level of objects and morphisms.

Assume (𝒰,𝔓)({\mathcal{U}},\mathfrak{P}) satisfies condition ()b(*)_{b} and the following condition

  • ()b(**)_{b}

    if γ\gamma is a bb-path in (Ui,di)(U_{i},d_{i}) and γ\gamma^{\prime} is a bb-path in (Uj,dj)(U_{j},d_{j}) such that γbγ\gamma\sim_{b}\gamma^{\prime} in (X,d)(X,d) then there exists a bb-path δ\delta in UijU_{ij} which is the same bb-map as γ\gamma in UiU_{i} with the metric did_{i} and as γ\gamma^{\prime} in UjU_{j} with the metric djd_{j}.

Then the morphism (12)(\ref{eq:comparisoncolim}) is an isomorphism of groupoids. In other words, if (𝒰,𝔓)({\mathcal{U}},\mathfrak{P}) satisfies conditions ()b(*)_{b} and ()b(**)_{b} then MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) is the colimit of the system of groupoids {MDπ1b(Ui1,,ir,𝔓)}i1.ir\{MD\pi_{1}^{b}(U_{i_{1},...,i_{r}},\mathfrak{P})\}_{i_{1}....i_{r}} (this is the usual formulation of Seifert-van Kampen Theorem for groupoids).

Proof of Theorem 55.

Since 𝒰{\mathcal{U}} is a subanalytic cover the mapping σ\sigma is always surjective at the level of objects.

Let γ\gamma be a bb-path in MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}). As in the proof of Theorem 54 we write [γ][\gamma] as

[γ]=[γ0δ10][δ11γ1δ20][δr1αr],[\gamma]=[\gamma_{0}\cdot\delta_{1}^{0}]\cdot[\overleftarrow{\delta_{1}^{1}}\cdot\gamma_{1}\cdot\delta_{2}^{0}]\cdot...\cdot[\overleftarrow{\delta_{r}^{1}}\cdot\alpha_{r}],

where each factor is a bb-path in MDπ1b(Ui,𝔓)MD\pi_{1}^{b}(U_{i},\mathfrak{P}) for a certain ii. This shows surjectivity at the level of morphisms.

By theorem 54, in order to prove the isomorphism of groupoids under condition ()b(**)_{b} it is enough to show that the groupoid L:=colim{MDπ1b(Ui1,,ir,𝔓)}i1.irL:=\mathrm{colim}\{MD\pi_{1}^{b}(U_{i_{1},...,i_{r}},\mathfrak{P})\}_{i_{1}....i_{r}} and the natural morphisms bi:MDπ1b(Ui,𝔓)Lb_{i}:MD\pi_{1}^{b}(U_{i},\mathfrak{P})\to L satisfy:

  1. (i)

    Property ()b(\dagger)_{b}, and

  2. (ii)

    for any other groupoid KK and groupoid morphisms κi:MDπ1b(Ui,𝔓)K\kappa_{i}:MD\pi_{1}^{b}(U_{i},\mathfrak{P})\to K with the property ()b(\dagger)_{b} there exists a unique groupoid morphism κ:LK\kappa:L\to K such that κbi=κi\kappa{\circ}b_{i}=\kappa_{i} for any iIi\in I.

Both assertions follow from the definition of colimit.

Now,we particularize for the MD fundamental group, similarly to the classical Seifert-van Kampen Theorem:

Corollary 56.

Let 𝒰={U1,U2}iI{\mathcal{U}}=\{U_{1},U_{2}\}_{i\in I} be a finite cover of a metric subanalytic germ (X,0,d)(X,0,d) by subanalytic open subsets UiU_{i} with metrics did_{i} and djd_{j}, both of type (a) or (b). Assume that U1U_{1}, U2U_{2} and U1U2U_{1}\cap U_{2} are bb-connected. Choose a bb-point 𝔭\mathfrak{p} in U1U2U_{1}\cap U_{2}. If 𝒰{\mathcal{U}} satisfies property ()b(**)_{b} then MDπ1b(X,𝔭)MD\pi_{1}^{b}(X,\mathfrak{p}) is isomorphic to the amalgamated product

MDπ1b(U1,𝔭)MDπ1b(U1U2,𝔭)MDπ1b(U2,𝔭).MD\pi_{1}^{b}(U_{1},\mathfrak{p})*_{MD\pi_{1}^{b}(U_{1}\cap U_{2},\mathfrak{p})}MD\pi_{1}^{b}(U_{2},\mathfrak{p}).
Proof.

We have to check that Property ()b(*)_{b} holds. Let 𝔭1\mathfrak{p}_{1}, 𝔭2\mathfrak{p}_{2} be bb-points in U1U_{1} and U2U_{2} respectively. If 𝔭1b𝔭2\mathfrak{p}_{1}\sim_{b}\mathfrak{p}_{2} then by property ()b(**)_{b} there is a point 𝔮\mathfrak{q} in U1U2U_{1}\cap U_{2} which is bb-equivalent to both. Choose a path γ\gamma connecting 𝔮\mathfrak{q} with 𝔭\mathfrak{p} (use that U1U2U_{1}\cap U_{2} is bb-path connected). We can choose γ\gamma for both required paths in condition ()b(*)_{b} and choose the homotopies to be constant and equal to γ\gamma. ∎

Corollary 57.

Let (X,0,d)(X,0,d) be a metric subanalytic germ. Fix bb and let 𝒰{\mathcal{U}} be a cover satisfying Condition ()b(**)_{b}. Assume that each intersection of up to 33 open subsets of the cover is bb-connected and has trivial MD bb-fundamental group. Then the MD bb-fundamental group of XX is isomorphic to the topological fundamental group of the nerve of the cover.

Proof.

Choose a set 𝔓\mathfrak{P} of base points having one point at each intersection of 11, 22 or 33 open subsets of the cover. By the bb-connectivity assumptions condition ()b(*)_{b} is trivially satisfied. Then by Theorem 55 MDπ1b(X,0,d,𝔓)MD\pi_{1}^{b}(X,0,d,\mathfrak{P}) is the coproduct of the system of fundamental groupoids associated to the cover. From this point the proof is reduced to the same statement in the topological category (see [9] Proposition 4.G2).

5. Comparison theorems

5.1. The \infty-MD homotopy sets and the homotopy of the link

Proposition 58.

Let (X,Y,x0,d,𝔭)(X,Y,x_{0},d,\mathfrak{p}) be a pointed pair of metric subanalytic germs. Fix ϵ>0\epsilon>0 small enough so that (LX,ϵ,LY,ϵ,𝔭ϵ):=(X,Y,Im(𝔭)𝕊ϵ)(L_{X,\epsilon},L_{Y,\epsilon},\mathfrak{p}_{\epsilon}):=(X,Y,\mathrm{Im}(\mathfrak{p})\cap{\mathbb{S}}_{\epsilon}) is the link of the pair. There is a bijection

ζ:πn(LX,ϵ,LY,ϵ,𝔭ϵ)MDπn(X,Y,𝔭)\zeta:\pi_{n}(L_{X,\epsilon},L_{Y,\epsilon},\mathfrak{p}_{\epsilon})\xrightarrow{\sim}MD\pi_{n}^{\infty}(X,Y,\mathfrak{p})

where πn(LX,ϵ,LY,ϵ,𝔭ϵ)\pi_{n}(L_{X,\epsilon},L_{Y,\epsilon},\mathfrak{p}_{\epsilon}) denotes the standard nn-th homotopy set. The bijection is a group isomorphism whenever a group structure is defined.

Proof.

We can assume that 𝔭\mathfrak{p} is a normal point by independence of the base point (see Proposition 40). Then 𝔭ϵ=𝔭(ϵ)\mathfrak{p}_{\epsilon}=\mathfrak{p}(\epsilon).

It is well known that πn(LX,ϵ,LY,ϵ,𝔭ϵ)\pi_{n}(L_{X,\epsilon},L_{Y,\epsilon},\mathfrak{p}_{\epsilon}) is the quotient of the set of subanalytic nn-loops in (LX,ϵ,LY,ϵ,𝔭ϵ)(L_{X,\epsilon},L_{Y,\epsilon},\mathfrak{p}_{\epsilon}) by subanalytic homotopies (this can be done using a subanalytic triangulation and the simplicial approximation theorem). Now to define ζ\zeta, we choose h:(C(LX,ϵ,LY,ϵ,𝔭ϵ),0)(X,Y,Im(𝔭),x0)h:(C(L_{X,\epsilon},L_{Y,\epsilon},\mathfrak{p}_{\epsilon}),0)\xrightarrow{\sim}(X,Y,\text{Im}(\mathfrak{p}),x_{0}) a subanalytic homeomorphism germ that gives conical structure (see Definition 8 and Remark 9) to (X,0)(X,0) compatible with the subgerm Im(𝔭)Im(\mathfrak{p}) and such that h(x,t)=t||h(x,t)||=t. Then, given a subanalytic nn-loop ψ:In(LX,ϵ,LY,ϵ,𝔭ϵ)\psi:I^{n}\to(L_{X,\epsilon},L_{Y,\epsilon},\mathfrak{p}_{\epsilon}) we define ζ(ψ):C(In)X\zeta(\psi):C(I^{n})\to X by the formula ζ(ψ)(x,t)=h(ψ(x),t)\zeta(\psi)(x,t)=h(\psi(x),t). The assignment is obviously well defined.

Let us show surjectivity. Let φ:C(In)X\varphi:C(I^{n})\to X be any \infty-MD nn-loop in MDπn(X,Y,𝔭)MD\pi_{n}^{\infty}(X,Y,\mathfrak{p}). Let us denote h1(x)=(α(x),τ(x))h^{-1}(x)=(\alpha(x),\tau(x)), where hh is the homeomorphism we chose above and we are taking coordinates (y,t)(y,t) in C(LX,ϵ)C(L_{X,\epsilon}) as in Notation 5. We see first that φ\varphi is bb-homotopic to (,b)(\infty,b)-loop φ\varphi^{\prime} satisfying τ(φ(y,t))=t\tau^{\prime}(\varphi^{\prime}(y,t))=t. Let ry(t)r_{y}(t) be the inverse map germ of the map germ τ(φ(y,_))\tau(\varphi(y,\_)) defined on (0,ϵ)(0,\epsilon). Then it is clear that τ(φ(y,ry(t)))=t\tau(\varphi(y,r_{y}(t)))=t. We can take φ(y,t):=φ(y,ry(t))\varphi^{\prime}(y,t):=\varphi(y,r_{y}(t)) since it is clear that H1:C(In+1)(X,Y)H_{1}:C(I^{n+1})\to(X,Y) defined by H1(y,s,t):=φ(y,(1s)t+sry(t))H_{1}(y,s,t):=\varphi(y,(1-s)t+sr_{y}(t)) defines a homotopy from φ\varphi to φ\varphi^{\prime}.

Now we see that φ\varphi^{\prime} is homotopic to φ′′(y,t):=h(αφ(y,ϵ),t)\varphi^{\prime\prime}(y,t):=h(\alpha\circ\varphi^{\prime}(y,\epsilon),t) which is obviously in the image of ζ\zeta. A homotopy from φ′′\varphi^{\prime\prime} to φ\varphi^{\prime} is H(y,s,t):=h(α(φ(y,ρ(t,s)),t)H(y,s,t):=h(\alpha(\varphi^{\prime}(y,\rho(t,s)),t) with t(0,ϵ)t\in(0,\epsilon), where ρ(t,s):=ϵ\rho(t,s):=\epsilon if t(1s)ϵt\geq(1-s)\epsilon and ρ(t,s):=t+(1s)ϵ\rho(t,s):=t+(1-s)\epsilon if t(1s)ϵt\leq(1-s)\epsilon.

Injectivity is proven applying the same procedure to the homotopies. ∎

5.2. Comparison of germs with the outer metric and their horn neighbourhoods

Definition 59.

Let ded_{e} denote the euclidean metric in n{\mathbb{R}}^{n}. Let (X,Y,de,O,𝔭)(X,Y,d_{e},O,\mathfrak{p}) be a pointed pair metric of subanalytic germs embedded in n{\mathbb{R}}^{n} with the outer metric ded_{e}. For b<b<\infty we denote by 𝒩bMDΓkb(X,Y,𝔭){\mathcal{N}}_{b}MD\Gamma^{b}_{k}(X,Y,\mathfrak{p}) the set of weak bb-maps φ:C(Ik)(n,O)\varphi:C(I^{k})\to({\mathbb{R}}^{n},O) satisfying

  1. (a)

    limt0max{de(φ(ty,t),X):yIk}tb=0.\lim_{t\to 0}\frac{max\{d_{e}(\varphi(ty,t),X):y\in I^{k}\}}{t^{b}}=0.

  2. (b)

    for any point 𝔮\mathfrak{q} in C(Ik)C(\partial I^{k}), the point φ𝔮\varphi{\circ}\mathfrak{q} is bb-equivalent to a point 𝔮:(0,ϵ)Y\mathfrak{q}^{\prime}:(0,\epsilon)\to Y.

  3. (c)

    for any normal point 𝔮\mathfrak{q} in C(Ik(Ik1×{1}))C(\partial I^{k}\setminus(I^{k-1}\times\{1\})), the point φq\varphi{\circ}q is bb-equivalent to 𝔭\mathfrak{p}.

We denote by 𝒩bMDπkb(X,Y,𝔭){\mathcal{N}}_{b}MD\pi^{b}_{k}(X,Y,\mathfrak{p}) the quotient of 𝒩MDΓkb(X,Y,𝔭){\mathcal{N}}MD\Gamma^{b}_{k}(X,Y,\mathfrak{p}) by weak bb-homotopies H:C(Ik+1)IH:C(I^{k+1})\to I, relative to C(Ik(Ik1×{1}))C(\partial I^{k}\setminus(I^{k-1}\times\{1\})), preserving the inclusion of In\partial I^{n} into YY and satisfying

limt0max{de(H(ty,t),X):yIk+1}tb=0.\lim_{t\to 0}\frac{max\{d_{e}(H(ty,t),X):y\in I^{k+1}\}}{t^{b}}=0.

The sets 𝒩bMDπkb(X,Y,𝔭){\mathcal{N}}_{b}MD\pi^{b}_{k}(X,Y,\mathfrak{p}) have a group structure whose product is the concatenation of bb-loops whenever this concatenation is possible.

Proposition 60.

With the notations of the previous definition there is a bijection (a group isomorphism when the group structure is defined)

ρ:𝒩bMDπkb(X,Y,𝔭)MDπkb(X,Y,𝔭)\rho:{\mathcal{N}}_{b}MD\pi^{b}_{k}(X,Y,\mathfrak{p})\to MD\pi^{b}_{k}(X,Y,\mathfrak{p})

for any kk and for any bb.

Proof.

As a consequence of Proposition 32 we can replace XX by its closure without loosing generality. So we assume that XX is closed.

Apply Lemma 35 to the pair BXB\supset X. Let 𝒮:={Si}iI{\mathcal{S}}:=\{S_{i}\}_{i\in I} and {fi}iI\{f_{i}\}_{i\in I} be the partition and the subanalytic maps predicted in the Lemma 35 . Then {S¯i}iI\{\overline{S}_{i}\}_{i\in I} is a closed cover of BB. Let UU be the union of the interiors of the sets SiS_{i}. Then UU is a dense subanalytic subset of BB. A straightforward adaptation of Proposition 32 shows the bijection

ι:𝒩bMDπkb,U(X,Y,𝔭)𝒩bMDπkb(X,Y,𝔭),\iota:{\mathcal{N}}_{b}MD\pi^{b,U}_{k}(X,Y,\mathfrak{p})\to{\mathcal{N}}_{b}MD\pi^{b}_{k}(X,Y,\mathfrak{p}),

where 𝒩bMDπkb,U(X,Y,𝔭){\mathcal{N}}_{b}MD\pi^{b,U}_{k}(X,Y,\mathfrak{p}) is defined in analogy with Definition 31.

Given any loop φ𝒩bMDΓkb,U(X,Y,𝔭)\varphi\in{\mathcal{N}}_{b}MD\Gamma^{b,U}_{k}(X,Y,\mathfrak{p}) we choose a representative {(Cj,gj)}jJ\{(C_{j},g_{j})\}_{j\in J} such that g(Cj)g(C_{j}) is contained in Si(j)S_{i(j)} for a certain i(j)i(j). Define ρ(σ):={(Cj,fi(j)gj)}jJ\rho^{\prime}(\sigma):=\{(C_{j},f_{i(j)}{\circ}g_{j})\}_{j\in J}. A straightforward application of the triangle inequality shows that ρ(σ)\rho^{\prime}(\sigma) is a well defined element in MDΓkb(X,Y,𝔭)MD\Gamma^{b}_{k}(X,Y,\mathfrak{p}). Applying the same procedure to the homotopies we get a well defined mapping

ρ:𝒩bMDπkb,U(X,Y,𝔭)MDπkb(X,Y,𝔭).\rho^{\prime}:{\mathcal{N}}_{b}MD\pi^{b,U}_{k}(X,Y,\mathfrak{p})\to MD\pi^{b}_{k}(X,Y,\mathfrak{p}).

Define ρ:=ρι1\rho:=\rho^{\prime}{\circ}\iota^{-1}. The mapping

MDπkb(X,Y,𝔭)𝒩bMDπkb(X,Y,𝔭)MD\pi^{b}_{k}(X,Y,\mathfrak{p})\to{\mathcal{N}}_{b}MD\pi^{b}_{k}(X,Y,\mathfrak{p})

induced by the inclusion is clearly a left-inverse of ρ\rho. It is also a right inverse because given φ𝒩bMDΓkb,U(X,Y,𝔭)\varphi\in{\mathcal{N}}_{b}MD\Gamma^{b,U}_{k}(X,Y,\mathfrak{p}), the weak bb-maps φ\varphi and ρφ\rho^{\prime}{\circ}\varphi are bb equivalent due to Condition (a) of Definition 59. ∎

For the next lemma recall the definition of horn neighborhood from [8]:

Definition 61 (Horn Neighborhood).

Let XX be a subanalytic germ embedded in n{\mathbb{R}}^{n}. We assume the vertex of the cone to be the origin in n{\mathbb{R}}^{n}. Let b+b\in{\mathbb{R}}^{+}. The bb-horn neighborhood of amplitude η\eta of XX in n{\mathbb{R}}^{n} is the union

b,η(X):=xXB(x,ηxb).{\mathcal{H}}_{b,\eta}(X):=\bigcup_{x\in X}B(x,\eta\|x\|^{b}).
Lemma 62.

With the same notations as above, for any fixed η>0\eta>0, the set 𝒩bMDπkb(X,Y,𝔭){\mathcal{N}}_{b}MD\pi^{b}_{k}(X,Y,\mathfrak{p}) is in a bijection with the direct limit of sets

limb>bMDπkb(b,η(X),b,η(Y),𝔭,de).\lim_{\stackrel{{\scriptstyle\longrightarrow}}{{b^{\prime}>b}}}MD\pi^{b^{\prime}}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p},d_{e}).
Proof.

Since we have the inclusion b1,η(X)b2,η(X){\mathcal{H}}_{b_{1},\eta}(X)\subset{\mathcal{H}}_{b_{2},\eta}(X) if b1>b2b_{1}>b_{2} we have natural maps MDπkb1(b1,η(X),b1,η(Y),𝔭)MDπkb2(b2,η(X),b2,η(Y),𝔭)MD\pi^{b_{1}}_{k}({\mathcal{H}}_{b_{1},\eta}(X),{\mathcal{H}}_{b_{1},\eta}(Y),\mathfrak{p})\to MD\pi^{b_{2}}_{k}({\mathcal{H}}_{b_{2},\eta}(X),{\mathcal{H}}_{b_{2},\eta}(Y),\mathfrak{p}) defined to be the composition of

MDπkb1(b1,η(X),b1,η(Y),𝔭)MDπkb1(b2,η(X),b2,η(Y),𝔭)MDπkb2(b2,η(X),b2,η(Y),𝔭).MD\pi^{b_{1}}_{k}({\mathcal{H}}_{b_{1},\eta}(X),{\mathcal{H}}_{b_{1},\eta}(Y),\mathfrak{p})\to MD\pi^{b_{1}}_{k}({\mathcal{H}}_{b_{2},\eta}(X),{\mathcal{H}}_{b_{2},\eta}(Y),\mathfrak{p})\to MD\pi^{b_{2}}_{k}({\mathcal{H}}_{b_{2},\eta}(X),{\mathcal{H}}_{b_{2},\eta}(Y),\mathfrak{p}).

This forms the direct system.

For any b>bb^{\prime}>b any nn-loop representing an element in MDπkb(b,η(X),b,η(Y),𝔭)MD\pi^{b^{\prime}}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p}), and the bb^{\prime}-homotopies connecting them satisfy property (a)(a) of Definition 59. So we have a homomorphism

ξ:limb>bMDπkb(b,η(X),b,η(Y),𝔭)𝒩bMDπkb(X,Y,𝔭).\xi:\lim_{\stackrel{{\scriptstyle\longrightarrow}}{{b^{\prime}>b}}}MD\pi^{b^{\prime}}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p})\to{\mathcal{N}}_{b}MD\pi^{b}_{k}(X,Y,\mathfrak{p}).

For any loop φ={(Cj,gj)}jJ\varphi=\{(C_{j},g_{j})\}_{j\in J} representing an element in 𝒩bMDπkb(X,Y,𝔭){\mathcal{N}}_{b}MD\pi^{b}_{k}(X,Y,\mathfrak{p}) we expand

maxjJ{max{de(gj(ty,t),X):(y,t)Cj}}=a1tb1+θ1(t),\max_{j\in J}\{\max\{d_{e}(g_{j}(ty,t),X):(y,t)\in C_{j}\}\}=a_{1}t^{b^{\prime}_{1}}+\theta_{1}(t),
maxj,jJ{max{de(gj(ty,t),gj(ty,t)):(y,t)CjCj}}=a2tb2+θ2(t),\max_{j,j^{\prime}\in J}\{\max\{d_{e}(g_{j}(ty,t),g_{j^{\prime}}(ty,t)):(y,t)\in C_{j}\cap C_{j^{\prime}}\}\}=a_{2}t^{b^{\prime}_{2}}+\theta_{2}(t),
maxjJ{max{de(gj(ty,t),Y):(y,t)CjC(Ik)}}=a3tb3+θ3(t),\max_{j\in J}\{\max\{d_{e}(g_{j}(ty,t),Y):(y,t)\in C_{j}\cap\partial C(I^{k})\}\}=a_{3}t^{b^{\prime}_{3}}+\theta_{3}(t),

where θi(t)\theta_{i}(t) vanishes at 0 to order higher than bib^{\prime}_{i}, and bi>bb^{\prime}_{i}>b for i=1,2,3i=1,2,3. Then φ\varphi represents an element in MDπkb′′(b′′,η(X),Y,𝔭)MD\pi^{b^{\prime\prime}}_{k}({\mathcal{H}}_{b^{\prime\prime},\eta}(X),Y,\mathfrak{p}) for any b′′b^{\prime\prime} satisfying b<b′′<min{b1,b2,b3}b<b^{\prime\prime}<\min\{b^{\prime}_{1},b^{\prime}_{2},b^{\prime}_{3}\}. This shows the surjectivity of ξ\xi. For the injectivity we apply the same argument to the homotopies. ∎

Proposition 63.

With the same notations as above, there is a bijection

limb>bMDπk(b,η(X),b,η(Y),𝔭,de)limb>bMDπkb(b,η(X),b,η(Y),𝔭,de).\lim_{\stackrel{{\scriptstyle\longrightarrow}}{{b^{\prime}>b}}}MD\pi^{\infty}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p},d_{e})\cong\lim_{\stackrel{{\scriptstyle\longrightarrow}}{{b^{\prime}>b}}}MD\pi^{b^{\prime}}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p},d_{e}).

The main part of the proof consists in given a weak bb^{\prime}-loop on the right hand side limit, finding a bb^{\prime}-equivalent continuous one on the left hand side. We perform an interpolation trick for that. We need some preliminary work.

Lemma 64.

Consider b>1b^{\prime}>1. Let 𝔭1,,𝔭l:(0,ϵ)˙b,η(X)\mathfrak{p}_{1},...,\mathfrak{p}_{l}:(0,\epsilon)\to\dot{\mathcal{H}}_{b^{\prime},\eta}(X) be bb^{\prime}-equivalent points. A point of the form iλi𝔭i\sum_{i}\lambda_{i}\mathfrak{p}_{i} with iλi=1\sum_{i}\lambda_{i}=1 is contained in ˙b′′,η(X)\dot{\mathcal{H}}_{b^{\prime\prime},\eta}(X) for every b′′<bb^{\prime\prime}<b and it is bb^{\prime}-equivalent to the 𝔭i\mathfrak{p}_{i}.

Proof.

Since the points are bb equivalent there exists a positive AA such that pi(t)=At+θi(t)||p_{i}(t)||=At+\theta_{i}(t), with limt0θi(t)/tb=0\lim_{t\to 0}\theta_{i}(t)/t^{b^{\prime}}=0 for all ii, and fixed λi\lambda_{i} such that iλi=1\sum_{i}\lambda_{i}=1 we have iλipi(t)=At+θ(t)||\sum_{i}\lambda_{i}p_{i}(t)||=At+\theta(t) with limt0θ(t)/tb=0\lim_{t\to 0}\theta(t)/t^{b^{\prime}}=0.

d(X,i=1lλi𝔭i(t))d(X,𝔭1(t))+d(𝔭1(t),i=1lλi𝔭i(t))d(X,\sum_{i=1}^{l}\lambda_{i}\mathfrak{p}_{i}(t))\leq d(X,\mathfrak{p}_{1}(t))+d(\mathfrak{p}_{1}(t),\sum_{i=1}^{l}\lambda_{i}\mathfrak{p}_{i}(t))\leq
η||𝔭1(x)||b+λid(𝔭1(t),𝔭i(t))η||λi𝔭i(x)||b+||ρ(t)||+λid(𝔭1(t),𝔭i(t),\eta||\mathfrak{p}_{1}(x)||^{b}+\sum\lambda_{i}d(\mathfrak{p}_{1}(t),\mathfrak{p}_{i}(t))\leq\eta||\sum\lambda_{i}\mathfrak{p}_{i}(x)||^{b}+||\rho(t)||+\sum\lambda_{i}d(\mathfrak{p}_{1}(t),\mathfrak{p}_{i}(t),

where limt0ρ(t)/tb=0\lim_{t\to 0}\rho(t)/t^{b^{\prime}}=0.

The above inequality implies that iλi𝔭i\sum_{i}\lambda_{i}\mathfrak{p}_{i} is contained in ˙b,′′η(X)\dot{\mathcal{H}}_{b{{}^{\prime\prime}},\eta}(X) for every b′′<bb^{\prime\prime}<b^{\prime}. ∎

Construction: the skeleton thickening decomposition.

For a set of points AA we denote by [A][A] its convex hull. A nn-simplex in N{\mathbb{R}}^{N} is the convex hull of n+1n+1 affinely independent points. More generally a polytope is the convex hull of finitely many points. A piecewise linear simplicial complex |K||K| is a locally finite union of simplexes such that any finite intersection of simplexes is a simplex. A simplex is maximal if it is not contained in any other different simplex. Now we decompose |K||K| as a union of subsets which refines the decomposition of |K||K| into maximal simplexes.

We start defining the decomposition for a single simplex. Let TT be the set of vertices of a kk dimensional simplex. For any subsets ffTf\subseteq f^{\prime}\subseteq T we define

bfthe baricenter of[f]b_{f}\ \text{the\ baricenter\ of}\ [f]
vf:=12bf+12v for any vfv^{f}:=\frac{1}{2}b_{f}+\frac{1}{2}v\ \text{\ \ for\ any }\ v\in f
ff:={vf:vf}for  any ffTf^{f^{\prime}}:=\{v^{f^{\prime}}:v\in f\}\ \text{for\ any }\ f\subseteq f^{\prime}\subseteq T
ff:=f′′:ff′′fff′′f^{*f^{\prime}}:=\bigcup_{f^{\prime\prime}:f\subseteq f^{\prime\prime}\subseteq f^{\prime}}f^{f^{\prime\prime}}
Refer to caption
Figure 5. Partial decomposition of the simplex T={v0,v1,v2}T=\{v_{0},v_{1},v_{2}\} where the edges are e0=[v0,v1]e_{0}=[v_{0},v_{1}], e1=[v1,v2]e_{1}=[v_{1},v_{2}] and e2=[v2,v0]e_{2}=[v_{2},v_{0}].

Note that [ff][f^{f^{\prime}}] is a simplex of the same dimension as [f][f], strictly contained in the interior of [f][f^{\prime}] and with faces parallel to the ones of [f][f] (see Figure 5). In particular [TT][T^{T}] is a simplex of dimension kk completely contained in [T][T].

Observe that, given fff\subset f^{\prime}, the sets [ff][f^{*f^{\prime}}] are of the same dimension than [f][f^{\prime}]. There are 2ss2^{s^{\prime}-s} possible subsets f′′f^{\prime\prime} with ff′′ff\subseteq f^{\prime\prime}\subseteq f^{\prime} if s=dim([f])s^{\prime}=dim([f^{\prime}]) and s=dim([f])s=dim([f]).

We have a decomposition

(13) [T]=[TT]f:fT[fT][T]=[T^{T}]\cup\bigcup_{f:f\subsetneq T}[f^{*T}]

where ff runs over the set of faces of TT of dimension less than kk. It can be checked that the sets in the decomposition are polytopes of dimension kk and only intersect along polytopes (faces) of smaller dimension.

Let us also see that every [fT][f^{*T}] is fibred over [ff][f^{f}] with fiber a cube [0,1]ks[0,1]^{k-s} with s=dim([f])s=dim([f]). Consider in [T][T] the pushforward metric of the euclidean metric in the standard kk-simplex [(1,,0),.,(0,,1)]k+1[(1,...,0),....,(0,...,1)]\subset{\mathbb{R}}^{k+1} by the mapping given by the barycentric coordinates in [T][T]. Consider the orthogonal projection π\pi from [T][T] to [f][f]. It can be easily checked by elementary geometry, that the restriction of π\pi to [fT][f^{T}] is a bijection onto [ff][f^{f}], and that π\pi restricted to [fT][f^{*T}] is a trivial bundle with base [ff][f^{f}] (or [fT][f^{T}]) and fiber a cube of the corresponding dimension. So, we have orthogonal trivializations

Γf,T:[fT][fT]×[0,1]ks\Gamma_{f,T}:[f^{*T}]\to[f^{T}]\times[0,1]^{k-s}

where the first projection to [fT][f^{T}] coincides with the orthogonal projection in [T][T] to [f][f] with respect to the mentioned metric. Note that by construction if ffTf\subseteq f^{\prime}\subseteq T and x[fT][fT]x\in[f^{*T}]\cap[{f^{\prime}}^{*T}] then the first projections of Γf,T(x)\Gamma_{f,T}(x) and Γf,T(x)\Gamma_{f^{\prime},T}(x) to [fT][f^{T}] and [fT][{f^{\prime}}^{T}] coincide.

We call the decomposition (13) the skeleton thickening decomposition of [T][T], and the system of mappings Γf,T\Gamma_{f,T} is the associated system of trivializations.

Let [T][T] be a kk-simplex and let [T][T^{\prime}] be a ll-dimensional face of [T][T]. The intersection of the pieces of the skeleton thickening decomposition of [T][T] with [T][T^{\prime}] gives induces the skeleton thickening decomposition of [T][T^{\prime}]. The associated system of trivializations of the skeleton thickening decomposition of [T][T^{\prime}] coincide with the relevant trivializations of [T][T], forgetting a [0,1]kl[0,1]^{k-l} factor on the right hand side.

Once we have defined the decomposition for a simplex, we define the skeleton thickening decomposition of |K||K| decomposing every maximal simplex [T][T] as in Equation (13). We consider the corresponding trivialization mappings Γf,T\Gamma_{f,T}.

Finally, for each symplex [T][T] of |K||K| we consider the unique affine homeomorphism

(14) τT:[TT][T].\tau_{T}:[T^{T}]\to[T].
Proof of Proposition 63.

The natural maps MDπk(b,η(X),b,η(Y),𝔭,de)MDπkb(b,η(X),b,η(Y),𝔭,de)MD\pi^{\infty}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p},d_{e})\to MD\pi^{b^{\prime}}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p},d_{e}) induce a map

limb>bMDπk(b,η(X),b,η(Y),𝔭,de)limb>bMDπkb(b,η(X),b,η(Y),𝔭,de).\lim_{\stackrel{{\scriptstyle\longrightarrow}}{{b^{\prime}>b}}}MD\pi^{\infty}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p},d_{e})\to\lim_{\stackrel{{\scriptstyle\longrightarrow}}{{b^{\prime}>b}}}MD\pi^{b^{\prime}}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p},d_{e}).

We start showing its surjectivity. Consider a bb^{\prime}-map ϕ={(Ci,gi)}iI\phi=\{(C_{i},g_{i})\}_{i\in I} representing an element of MDπkb(b,η(X),b,η(Y),𝔭,de)MD\pi^{b^{\prime}}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p},d_{e}). We claim that for any b<b′′<bb<b^{\prime\prime}<b^{\prime} there exists a continuous subanalytic kk-loop ϕ:C(Ik)b′′,η(X)\phi^{\prime}:C(I^{k})\to{\mathcal{H}}_{b^{\prime\prime},\eta}(X) that is weak b′′b^{\prime\prime}-homotopic to ϕ\phi relative to C(IkIk1×{1})C(\partial I^{k}\setminus I^{k-1}\times\{1\}) by a homotopy preserving the inclusion of C(Ik)C(\partial I^{k}) into b′′,η(Y){\mathcal{H}}_{b^{\prime\prime},\eta}(Y). The claim obviously shows the surjectivity. Let us prove it now.

By refining we can assume that CiCjC_{i}\cap C_{j} have empty interior. By the subanalytic triangulation theorem and Remark 9 there exists a piecewise linear simplicial complex |K||K| such that |K|=Ik|K|=I^{k} and a subanalytic homeomorphism h:C(|K|)C(Ik)h:C(|K|)\to C(I^{k}) of the form h(x,t)=(h1(x,t),t)h(x,t)=(h_{1}(x,t),t) in coordinates of C(Ik)C(I^{k}), that gives a subanalytic triangulation on C(Ik)C(I^{k}) refining {Ci}\{C_{i}\}, the natural strata of IkI^{k} and Im(𝔭)Im(\mathfrak{p}).

We will prove the claim for ϕh\phi{\circ}h instead of ϕ\phi. The proof for ϕ\phi consists in composing the obtained continuous kk-loop and homotopy with h1h^{-1}. So, from now on we abuse notation and identify ϕ\phi with ϕh\phi{\circ}h, and assume that the decomposition of IkI^{k} given by the CiC_{i}’s coincides with the cone of the piecewise linear triangulation of IkI^{k} induced by |K||K|.

Consider the skeleton thickening decomposition of |K||K|, together with the trivialization mappings associated with the maximal simplexes and their faces.

Since we are assuming that the decomposition C(Ik)=iICiC(I^{k})=\cup_{i\in I}C_{i} is the cone of the decomposition of |K||K| into maximal simplexes, we have that the maximal simplexes are indexed by II. Given any iIi\in I we define

(15) Ri:=[Ti]fTiTjf[fTj].R_{i}:=[T_{i}]\cup\bigcup_{f\subset T_{i}}\bigcup_{T_{j}\supset f}[f^{*T_{j}}].

We define a a weak bb^{\prime}-map (C(Ri),g¯i)(C(R_{i}),\overline{g}_{i}) of ϕ\phi as follows:

(16) g¯i(x,t)={gi(τTi(x),t)if x[TiTi]gi(τTi(y),t)ifx[fTi]andΓf,Ti(x)=(y,u)gi(τTi(y¯),t)ifx[fTj]andΓf,Tj(x)=(y,u)\overline{g}_{i}(x,t)=\begin{cases}g_{i}(\tau_{T_{i}}(x),t)&\quad\text{if }x\in[T_{i}^{T_{i}}]\\ g_{i}(\tau_{T_{i}}(y),t)&\quad\text{if}\ x\in[f^{*T_{i}}]\ \text{and}\ \Gamma_{f,T_{i}}(x)=(y,u)\\ g_{i}(\tau_{T_{i}}(\overline{y}),t)&\quad\text{if}\ x\in[f^{*T_{j}}]\ \text{and}\ \Gamma_{f,T_{j}}(x)=(y,u)\\ \end{cases}

where y¯[fTi]\overline{y}\in[f^{T_{i}}] is the image of y[fTj]y\in[f^{T_{j}}] by the compositions of the indentifications of [fTj][f^{T_{j}}] with [ff][f^{f}] and [ff][f^{f}] with [fTi][f^{T_{i}}] via the orthogonal projections in [Tj][T_{j}] and [Ti][T_{i}] respectively. It is obvious that g¯i\overline{g}_{i} is continuous.

Notice that for every ss-dimensional face f[T]f\in[T] and for every jj such that TjT_{j} contains ff, we have the product structure

C(Γf,Tj):C([fTj])C([fTj])×[0,1]ks,C(\Gamma_{f,T_{j}}):C([f^{*T_{j}}])\to C([f^{T_{j}}])\times[0,1]^{k-s},

and that gig_{i} is constant at the fibres of the composition of C(Γf,Tj)C(\Gamma_{f,T_{j}}) with the first projection. This, together with the fact that {(Ti,gi)}iI\{(T_{i},g_{i})\}_{i\in I} is a weak bb-map implies that ϕ¯\overline{\phi} is a weak bb-map.

Note that {([Ti],g¯i|[Ti])}i\{([T_{i}],\overline{g}_{i}|_{[T_{i}]})\}_{i} is a representative of ϕ¯\overline{\phi}, and that {([Ti],g¯i|[Ti])}i\{([T_{i}],\overline{g}_{i}|_{[T_{i}]})\}_{i} is clearly bb^{\prime}-homotopic to ϕ\phi. Then, ϕ¯\overline{\phi} is bb^{\prime}-homotopic to ϕ\phi.

Choose a subanalytic partition of unity {ρi}\{\rho_{i}\} adapted to {R˙i}\{\dot{R}_{i}\} in IkI^{k}. Then, we define ϕ(x,t)=iρi(x)g¯i(x,t)\phi^{\prime}(x,t)=\sum_{i}\rho_{i}(x)\overline{g}_{i}(x,t). This map is continuous since every ρi(x)g¯i(x)\rho_{i}(x)\overline{g}_{i}(x) is continuous in IkI^{k}, its image is inside b′′,η(X){\mathcal{H}}_{b^{\prime\prime},\eta}(X) by Lemma 64, and it is bb^{\prime}-equivalent to ϕ¯\overline{\phi}. This shows the claim and proves surjectivity.

Injectivity is proved with a similar argument applied to the homotopies. ∎

Now we are ready for the main result in this section:

Theorem 65.

Let (X,Y,de,x0,𝔭)(X,Y,d_{e},x_{0},\mathfrak{p}) be a pointed pair of metric subanalytic subgerms of n{\mathbb{R}}^{n} with the outer metric. For k1k\geq 1 in the absolute case and k2k\geq 2 in the relative one the following assertion holds. For any b<b<\infty there exists a b0>bb_{0}>b such that for any b(b,b0)b^{\prime}\in(b,b_{0}) there exists a positive ϵb\epsilon_{b^{\prime}} such that we have isomorphisms

MDHkb(X,Y;)Hk(b,η(X)Bϵb{x0},b,η(Y)Bϵb{x0};),MDH^{b}_{k}(X,Y;{\mathbb{Z}})\cong H_{k}({\mathcal{H}}_{b^{\prime},\eta}(X)\cap B_{\epsilon_{b^{\prime}}}\setminus\{x_{0}\},{\mathcal{H}}_{b^{\prime},\eta}(Y)\cap B_{\epsilon_{b^{\prime}}}\setminus\{x_{0}\};{\mathbb{Z}}),
MDπkb(X,Y,𝔭)πk(b,η(X)Bϵb{x0},b,η(Y)Bϵb{x0},𝔭(t0)),MD\pi^{b}_{k}(X,Y,\mathfrak{p})\cong\pi_{k}({\mathcal{H}}_{b^{\prime},\eta}(X)\cap B_{\epsilon_{b^{\prime}}}\setminus\{x_{0}\},{\mathcal{H}}_{b^{\prime},\eta}(Y)\cap B_{\epsilon_{b^{\prime}}}\setminus\{x_{0}\},\mathfrak{p}(t_{0})),

for t0t_{0} small enough.

Proof.

The MD Homology statement has been proved in [8].

By Proposition 63, Lemma 62 and Proposition 60 we have the isomorphism

MDπkb(X,Y,𝔭)limb>bMDπk(b,η(X),b,η(Y),p,de).MD\pi^{b}_{k}(X,Y,\mathfrak{p})\cong\lim_{\stackrel{{\scriptstyle\longrightarrow}}{{b^{\prime}>b}}}MD\pi^{\infty}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),p,d_{e}).

By Proposition 58 and the conical structure theorem for any bb^{\prime} there exists ϵb\epsilon_{b^{\prime}} so that we have

MDπk(b,η(X),b,η(Y),𝔭)πk(b,η(X)Bϵb{x0},b,η(Y)Bϵb{x0},p(t0)).MD\pi^{\infty}_{k}({\mathcal{H}}_{b^{\prime},\eta}(X),{\mathcal{H}}_{b^{\prime},\eta}(Y),\mathfrak{p})\cong\pi_{k}({\mathcal{H}}_{b^{\prime},\eta}(X)\cap B_{\epsilon_{b^{\prime}}}\setminus\{x_{0}\},{\mathcal{H}}_{b^{\prime},\eta}(Y)\cap B_{\epsilon_{b^{\prime}}}\setminus\{x_{0}\},p(t_{0})).

The rest of the proof is an easy adaptation of the invertible cobordism techniques of the proof of Theorem 107 and Corollary 118 of [8]. ∎

Remark 66.

Note that the skeleton thickening decomposition can be used as in the proof of 63 to solve the following convex interpolation problem in a simplicial complex: Let KK be a finite simplicial complex of any dimension. Assume you have continuous functions {gTi}i\{g_{T_{i}}\}_{i}, with values in a convex set AA, where each gTig_{T_{i}} is defined over the simplex [TiTi][T_{i}^{T_{i}}] (see Figure 5 and the skeleton thickening decomposition) with TiT_{i} a maximal dimensional simplex of KK. Find a continuous function defined over the whole |K||K| that extends the family of functions {gTi}Ti\{g_{T_{i}}\}_{T_{i}}.

5.3. Finiteness properties for germs with the inner or outer metric and comparision with their tangent cones

After the comparison theorem is established, easy adaptations of the corresponding results for the MD-homology in [8] yield the following set of results:

Corollary 67.

Let (X,Y,d,x0,𝔭)(X,Y,d,x_{0},\mathfrak{p}) be a pointed pair of metric subanalytic germs with the inner or the outer metric. Assume k1k\geq 1 for absolute homotopy groups and k2k\geq 2 for relative ones. The following assertions hold

  1. (i)

    The groups MDπkb(X,Y,𝔭)MD\pi^{b}_{k}(X,Y,\mathfrak{p}) are finitely presented.

  2. (ii)

    There is a finite set of rational numbers {b1,,bn}\{b_{1},...,b_{n}\} such that if [b,b][b,b^{\prime}] does not intersect it then the natural homomorphism

    MDπkb(X,Y,𝔭)MDπkb(X,Y,𝔭)MD\pi^{b^{\prime}}_{k}(X,Y,\mathfrak{p})\to MD\pi^{b}_{k}(X,Y,\mathfrak{p})

    for any kk as above.

  3. (iii)

    The group MDπk1(X,dout,𝔭)MD\pi^{1}_{k}(X,d_{out},\mathfrak{p}) (for the outer metric) is isomorphic to the kk-th homotopy group of the punctured tangent cone of XX.

  4. (iv)

    The group MDπk1(X,dinn,𝔭)MD\pi^{1}_{k}(X,d_{inn},\mathfrak{p}) (for the inner metric) is isomorphic to the kk-th homotopy group of the punctured Gromov tangent cone of XX (see [1] for the definition).

5.4. Metric Hurewicz isomorphism theorem

After Theorem 65 we can prove now Theorem  47:

Proof of Theorem 47.

For the outer metric, given the previous theorem, the proof gets reduced to the routine checking that the comparision isomorphisms are compatible with the Hurewicz homomorphisms in each category. For the inner metric, by [2], there exists a different embedding of the germ such that XX is Lipschitz normally embedded for the new embedding, and the metric is Lipschitz equivalent with the original inner metric of XX. ∎

5.5. MD homotopy groups of bb-cones

We recall that given a bounded subanalytic set LkL\subset{\mathbb{R}}^{k} and b(0,+)b\in\cap(0,+\infty) we call the outer bb-cone (CL,outb,dout)(C_{L,out}^{b},d_{out}) over LL to the germ at the origin of the set

CL,outb={(xtb,t)k×;xL and t[0,+)}C_{L,out}^{b}=\{(xt^{b},t)\in{\mathbb{R}}^{k}\times{\mathbb{R}};\,x\in L\mbox{ and }t\in[0,+\infty)\}

with the outer metric doutd_{out}. Note that the link of (CL,outb,dout)(C_{L,out}^{b},d_{out}) with the induced metric is bilispchitz to (L,dout)(L,d_{out}). When b=1b=1, we have denoted CL,out1C_{L,out}^{1} simply by C(L)C(L) and call it the straight cone over LL.

If bb\in{\mathbb{Q}} the outer bb-cone is subanalytic, and if bb\notin{\mathbb{Q}} it is at least definable in the OO-minimal structure an{\mathbb{R}}_{an}^{{\mathbb{R}}}.

It is also convenient to have a definition of inner metric bb-cones. Let LkL\subset{\mathbb{R}}^{k} be a subanalytic set. The inner metric bb-cone of LL will be a straight cone together with a metric such that its link with the induced metric is bilipschitz equivalent to (L,dinn)(L,d_{inn}). If LL is Lipschitz normally embedded (that is, the inner and outer metrics are bi-Lipschitz equivalent), then the inner bb-cone is by definition (CL,innb,dinn):=(CL,outb,dout)(C_{L,inn}^{b},d_{inn}):=(C_{L,out}^{b},d_{out}). If LL is not Lipschitz normally embedded, by [2] there exists a different subanalytic embedding LLkL\approx L^{\prime}\hookrightarrow{\mathbb{R}}^{k^{\prime}} such that LL^{\prime} is Lipschitz normally embedded, and the metric of (L,dout)(L^{\prime},d_{out}) is Lipschitz equivalent with (L,dinn)(L,d_{inn}), the inner metric that LL inherits from its original embedding in k{\mathbb{R}}^{k}. Then we define (CL,innb,dinn):=(CL,outb,dout)(C_{L,inn}^{b},d_{inn}):=(C_{L^{\prime},out}^{b},d_{out}). The following remark shows the independence on the embedding and gives an intrinsic description of the inner cone.

Remark 68.

Let gg be the Riemannian metric in the smooth part of LL inducing the inner metric in LL. Then, up to bi-Lipschitz equivalence, the Riemannian metric (dt)2+tbg(dt)^{2}+t^{b}g induces the metric in CLb,innC_{L}^{b,inn}.

Proposition 69.

Given a bounded connected subanalytic set LkL\subset{\mathbb{R}}^{k} we have for any k1k\geq 1

  1. a)

    If b<bb^{\prime}<b, then MDπkb(CL,outb,𝔭)MD\pi_{k}^{b^{\prime}}(C_{L,out}^{b},\mathfrak{p}) and MDπkb(CL,innb,𝔭)MD\pi_{k}^{b^{\prime}}(C_{L,inn}^{b},\mathfrak{p}) are trivial.

  2. b)

    If bbb^{\prime}\geq b then we have isomorphisms

    MDπkb(CL,outb,𝔭)MDπkb(CL,innb,𝔭)πk(L,𝔭0)MD\pi_{k}^{b^{\prime}}(C_{L,out}^{b},\mathfrak{p})\cong MD\pi_{k}^{b^{\prime}}(C_{L,inn}^{b},\mathfrak{p})\cong\pi_{k}(L,\mathfrak{p}_{0})
Proof.

By definition of the inner cone it is enough to work for the outer cone. In that case the result is a corollary of Theorem 65. ∎

6. The MD-fundamental group and MD-Homology of normal surface singularities with the inner metric

Let (X,0,dinn)(X,0,d_{inn}) be a complex normal surface singularity with its inner metric. Its Lipschitz geometry is completely described in [3]. Before we summarize it we recall the definition of some special metric subanalytic germs for convenience of the reader. They can be read in Section 11 in [3].

Definition 70 ([3]).

Choose 1q<q1\leq q<q^{\prime}. Let D2D1D_{2}\supset D_{1} be the concentric discs centered at the origin of 2{\mathbb{R}}^{2} of radii 11 and 22 and denote by D˙1\dot{D}_{1} the open disc. Consider the subset of 3{\mathbb{R}}^{3} given by Cq(D2)Cq(D˙1)C^{q}(D_{2})\setminus C^{q^{\prime}}(\dot{D}_{1}) with the inner metric and denote by gg its riemanninan tensor. The germ A(q,q){A}^{*}(q,q^{\prime}) is equal to 𝕊1×((Cq(D2)Cq(D˙1)){O}){\mathbb{S}}^{1}\times((C^{q}(D_{2})\setminus C^{q^{\prime}}(\dot{D}_{1}))\setminus\{O\}) with the riemannian metric g+t2dθ2g+t^{2}d\theta^{2} where θ\theta parametrizes the factor 𝕊1{\mathbb{S}}^{1} and tt is a l.v.a. parameter of a conical structure on (Cq(D2)Cq(D1),O)(C^{q}(D_{2})\setminus C^{q^{\prime}}(D_{1}),O). We denote by A(q,q){A}(q,q^{\prime}) the completion of A(q,q){A}^{*}(q,q^{\prime}), which adds a point that we call the vertex; A(q,q){A}(q,q^{\prime}) is homeomorphic to the cone over the annulus D2D˙1D_{2}\setminus\dot{D}_{1}. The metric extends to the vertex.

Definition 71.

Let q+q\in{\mathbb{R}}^{+} and let ϕ:FF\phi:F\to F be an orientation-preserving subanalytic diffeomorphism of a compact oriented surface FF in n{\mathbb{R}}^{n}. Consider the inner qq-cone CF,innq(F)C_{F,inn}^{q}(F) of FF and let gg be its riemannian tensor. Consider the extension ϕ¯:CF,innq(F)CF,innq(F)\overline{\phi}:C_{F,inn}^{q}(F)\to C_{F,inn}^{q}(F) given by ϕ¯(xtq,t)=(ϕ(x)tq,t)\overline{\phi}(xt^{q},t)=(\phi(x)t^{q},t). We denote by B(F,ϕ,q){B}^{*}(F,\phi,q) the mapping torus of ϕ¯|CF,innq(F){O}\overline{\phi}|_{C_{F,inn}^{q}(F)\setminus\{O\}} with the metric given by g+t2dθ2g+t^{2}d\theta^{2} where θ\theta parametrizes the factor 𝕊1{\mathbb{S}}^{1} and tt is a l.v.a. parameter of a conical structure on CF,innq(F)C_{F,inn}^{q}(F). We denote by B(F,ϕ,q){B}(F,\phi,q) the completion of B(F,ϕ,q){B}^{*}(F,\phi,q), which adds a point that we call the vertex; B(F,ϕ,q){B}(F,\phi,q) is homeomorphic to the cone over the mapping torus of ϕ\phi. The metric extends to the vertex. Moreover there is a locally trivial fibration to the puncture disc given by the coordinates (t,θ)(t,\theta).

Due to [3] the inner Lipschitz geometry of (X,0,dinn)(X,0,d_{inn}) is described in the following terms: there is a finite number of rational numbers 1=q1<q2<<qn1=q_{1}<q_{2}<...<q_{n} and a canonical subanalytic decomposition in closed subspaces

(17) X=k=1nYk1i<jnZi,jX=\bigcup_{k=1}^{n}Y_{k}\cup\bigcup_{1\leq i<j\leq n}Z_{i,j}

that endowed with the inner metric have the following properties

  1. (1)

    Y1Y_{1} is metrically conical, that is bi-Lipschitz subanalytic homeomorphic to the 11-cone over its link with the inner metric.

  2. (2)

    There is a subanalytic map ξ:XY1¯Dϵ\xi:\overline{X\setminus Y_{1}}\to D_{\epsilon}^{*} which is a locally trivial fibration, where DϵD^{*}_{\epsilon} denotes the punctured disc of a certain radius ϵ\epsilon. The map restricts to a locally trivial fibration over each piece of the decomposition of XY1X\setminus Y_{1}.

  3. (3)

    For each i>1i>1 denote by ϕi:FiFi\phi_{i}:F_{i}\to F_{i} the monodromy of ξ:(Yi)ϵ|DϵDϵ\xi:(Y_{i})_{\epsilon}|_{\partial D^{*}_{\epsilon}}\to\partial D^{*}_{\epsilon}. Then YiY_{i} is bilipschitz subanalytic diffeomorphic to B(Fi,ϕi,qi)B(F_{i},\phi_{i},q_{i}) in a such a way that ξ\xi is compatible way with ξ\xi and the natural projection of B(F,ϕi,qi)B(F,\phi_{i},q_{i}) to its base puctured disc given by the coordinates (t,θ)(t,\theta).

  4. (4)

    Every Zi,jZ_{i,j} is bilipschitz subanalytic diffeomorphic to a (possibly empty) disjoint union of pieces A(qi,qj)A(q_{i},q_{j}).

  5. (5)

    For a sufficiently small ϵ>0\epsilon>0 the decomposition

    (18) Xϵ=k=1n(Yk)ϵ1i<jn(Zi,j)ϵX_{\epsilon}=\bigcup_{k=1}^{n}(Y_{k})_{\epsilon}\cup\bigcup_{1\leq i<j\leq n}(Z_{i,j})_{\epsilon}

    is a (non-minimal) JSJ-decomposition of the link XϵX_{\epsilon} (here ZϵZ_{\epsilon} denotes Z𝕊ϵZ\cap{\mathbb{S}}_{\epsilon}), and the decomposition (17) is the cone over this decomposition by a subanalytic conical structure. In [3] it is proved that this JSJ decomposition together with the rates qiq_{i} is canonically determined and determines the inner Lipschitz geometry of (X,0)(X,0). This decomposition does not coincide in general with the minimal JSJ-decomposition.

  6. (6)

    Any continuous path joining two points in adjacent pieces in the decomposition (18) passes through the common boundary.

Remark 72.

Although in [3] the authors do not mention the subanalyticity of the decompostition, it holds by their construction: they start with a subanalytic decomposition of (X,0)(X,0) built from a carrousel associated with the discriminant of a generic projection (see Section 12 of [3]), and after glueing finitely many pieces together in order to reach the canonical decomposition. Each of the pieces of the initial decomposition are subanalytic bilipschitz homeomorphic to the models described above. The same statement for the canonical decomposition holds since the gluing rules of Lemma 13.1 in [3] preserve subanalyticity because they consists in gluing a finite number of subanalytic sets.

Notice also that the monodromy of the subanalytic fibration ξ\xi and its restriction to each of the pieces has a subanalytic representative as an application of Hardt’s trivialization theorem.

For the MD fundamental group computation below one can work instead with the initial subanalytic decomposition of (X,0)(X,0) produced in Section 12 of [3] without any further gluing, obtaining the same result.

Observe that YkY_{k} meets ZijZ_{ij} if and only if k=ik=i or k=jk=j. We start also recalling that for every k{1,,n}k\in\{1,...,n\} we have that the boundary (Yk)ϵ(\partial Y_{k})_{\epsilon} is a disjoint union of tori. Each tori is fibred over 𝕊1{\mathbb{S}}^{1} by ξ\xi, with fibre a disjoint union of circles. The boundary Yk\partial Y_{k} is subanalytically homeomorphic to the cone over (Yk)ϵ(\partial Y_{k})_{\epsilon}. With respect to the induced inner metric in Yk\partial Y_{k}, the base 𝕊1{\mathbb{S}}^{1} of the fibration collapses its metric at rate 11 and the fibres at rate qkq_{k}. Choose a collar (Ck)ϵ(C_{k})_{\epsilon} of (Yk)ϵ(\partial Y_{k})_{\epsilon} in (Yk)ϵ(Y_{k})_{\epsilon}. This collar is the link of a piece CkYkC_{k}\subset Y_{k} with the property that CkC_{k} is a disjoint union of anular pieces of type A(qk,qk)A(q_{k},q_{k}). It will be important later that both rates of the annular piece are equal. The piece CkC_{k} decomposes as a disjoint union as

Ck=CkCk+,C_{k}=C_{k}^{-}\coprod C_{k}^{+},

where CkC_{k}^{-} contains the boundary components of YkY_{k} that intersect pieces of the decomposition of XX collapsing slower (that is YkZk1,k\partial Y_{k}\cap Z_{k-1,k}), and Ck+C_{k}^{+} contains the boundary components of YkY_{k} that intersect pieces collapsing faster (that is YkZk,k+1\partial Y_{k}\cap Z_{k,k+1}). Notice that C1C_{1}^{-} and Cn+C_{n}^{+} are empty. We choose the CkC_{k} so that ξ|Ck\xi|_{C_{k}} is a locally trivial fibration.

Define

Z~i,j:=Zi,jCjCi+\widetilde{Z}_{i,j}:=Z_{i,j}\cup C_{j}^{-}\cup C_{i}^{+}

The piece Z~i,j\widetilde{Z}_{i,j} is the union of Zi,jZ_{i,j} with all the constant collapsing rate annular pieces defined above that intersect it.

For every b[1,+)b\in[1,+\infty) we define

U>b:=qk>bYkqi>bZ~i,j.U_{>b}:=\bigcup_{q_{k}>b}Y_{k}\cup\bigcup_{q_{i}>b}\widetilde{Z}_{i,j}.
U~>b:=qk>bYkqj>bZ~i,j,\widetilde{U}_{>b}:=\bigcup_{q_{k}>b}Y_{k}\cup\bigcup_{q_{j}>b}\widetilde{Z}_{i,j},

For any b[1,+)b\in[1,+\infty) we define the cover 𝒰b{\mathcal{U}}^{b} of XX that we will use to apply our version of the Seifert van Kampen Theorem for the MDπ1bMD\pi^{b}_{1} of Theorem 55:

(19) 𝒰b:={Yk:qkb}{Z~i,j:qi<qjb}{U~>b.}{\mathcal{U}}_{b}:=\{Y_{k}:q_{k}\leq b\}\cup\{\widetilde{Z}_{i,j}:q_{i}<q_{j}\leq b\}\cup\{\widetilde{U}_{>b}.\}

Note that the intersection of any 3 sets in 𝒰b{\mathcal{U}}_{b} is empty. Whenever the intersection of two is non-empty it is a union of CkC_{k}^{-} and Ck+C_{k}^{+} pieces.

We fix a l.v.a. subanalytic conical structure

(20) h:C(Xϵ)(X,0)h:C(X_{\epsilon})\to(X,0)

compatible with the decomposition in (17) and with the CkC_{k}^{-} and Ck+C_{k}^{+} pieces. If QQ is a set of points in XϵX_{\epsilon} we consider the points given by retraction lines from points in QQ and denote

(21) 𝔥(Q):={h|C({z}):zQ}\mathfrak{h}(Q):=\{h|_{C(\{z\})}:z\in Q\}

where C({z})C(\{z\}) denotes the subcone of C(Xϵ)C(X_{\epsilon}).

For any b[1,+)b\in[1,+\infty) we choose a collection PbP^{b} of points that meets all the connected components of the interior of (U~>b)ϵ(\widetilde{U}_{>b})_{\epsilon}, of (Yk)ϵ(Y_{k})_{\epsilon}, (Ck)ϵ(C_{k}^{-})_{\epsilon} and (Ck+)ϵ(C_{k}^{+})_{\epsilon} for every kk with qkbq_{k}\leq b and of (Zi,j)ϵ(Z_{i,j})_{\epsilon} with qi<qjbq_{i}<q_{j}\leq b. The conical structure induces a collection l.v.a points 𝔓b:=𝔥(Pb)\mathfrak{P}^{b}:=\mathfrak{h}(P^{b}) in XX.

Proposition 73.

For any b[1,+)b\in[1,+\infty) the pair (𝒰b,𝔓b)({\mathcal{U}}^{b},\mathfrak{P}^{b}) satisfies conditions ()b(*)_{b} and ()b(**)_{b} in Theorem 55 considering in every set of 𝒰b{\mathcal{U}}^{b} the inner metric did_{i} induced by inner metric in (X,0)(X,0), that is as in case (b) at the beginning of Section 4.2. Consequently, the groupoid MDπ1b(X,0,dinn,𝔓b)MD\pi_{1}^{b}(X,0,d_{inn},\mathfrak{P}^{b}) is the colimit of the system of groupoids associated to the cover 𝒰b{\mathcal{U}}^{b}.

Proof.

Condition ()b(*)_{b} follows from ()b(**)_{b} since the cover only has 22-fold intersections. We check condition ()b(**)_{b}. Let CC be the intersection of two sets BB and BB^{\prime} of the decomposition. Then CC is either a CkC_{k}^{-} or a Ck+C_{k}^{+} piece, which we call CC. The collapsing rate of this annular piece is qk<bq_{k}<b. This implies that

dinn((BC)ϵ,(BC)ϵ)=aϵqk+θ(ϵ)d_{inn}((B\setminus C)_{\epsilon},(B^{\prime}\setminus C)_{\epsilon})=a\epsilon^{q_{k}}+\theta(\epsilon)

for a>0a>0 and θ(ϵ)\theta(\epsilon) decreasing faster than ϵqk\epsilon^{q_{k}}. Moreover, dinn((BC)ϵ,(XC)ϵ)d_{inn}((B\setminus C)_{\epsilon},(X\setminus C)_{\epsilon}) and dinn((BC)ϵ,(XC)ϵ)d_{inn}((B^{\prime}\setminus C)_{\epsilon},(X\setminus C^{\prime})_{\epsilon}) are of the same form.

Let γ:C(I)B\gamma:C(I)\to B and γ:C(I)B\gamma^{\prime}:C(I)\to B^{\prime} be weak bb-paths which are bb-equivalent in XX. Then for any fixed sIs\in I we have that dinn(γ(s,t),γ(s,t))d_{inn}(\gamma(s,t),\gamma^{\prime}(s,t)) decreases faster than tbt^{b} where b>qkb>q_{k}. This, together with property (6)(6) of the decomposition and the previous bound implies that either γ(s,t)\gamma(s,t) or γ(s,t)\gamma^{\prime}(s,t) belongs to CC for tt small enough. Therefore we split II as a union I=IγIγI=I_{\gamma}\cup I_{\gamma^{\prime}} such that IγI_{\gamma} and IγI_{\gamma^{\prime}} are disjoint unions of closed intervals and IγIγI_{\gamma}\cap I_{\gamma^{\prime}} is finite, γ|C(Iγ)\gamma|_{C(I_{\gamma})} has image in CC and γ|C(Iγ)\gamma^{\prime}|_{C(I_{\gamma^{\prime}})} has image in CC. The path δ:C(I)C\delta:C(I)\to C defined by δ|C(Iγ):=γ|C(Iγ)\delta|_{C(I_{\gamma})}:=\gamma|_{C(I_{\gamma})} and δ|C(Iγ):=γ|C(Iγ)\delta|_{C(I_{\gamma^{\prime}})}:=\gamma^{\prime}|_{C(I_{\gamma^{\prime}})} is bb-equivalent both to γ\gamma and γ\gamma^{\prime}. ∎

We study in the following lemmata the MDπ1bMD\pi^{b}_{1} of each element of the cover 𝒰b{\mathcal{U}}^{b} and their intersections considering always the inner metric on them induced by inner metric in (X,0)(X,0), that is as in case (b) at the beginning of Section 4.2.

Lemma 74.

There is an isomorphism of fundamental groupoids

MDπ1b(Yk,𝔥(Q))π1((Yk)ϵ,Q)for bqk,MD\pi^{b}_{1}(Y_{k},\mathfrak{h}(Q))\cong\pi_{1}((Y_{k})_{\epsilon},Q)\ \ \text{for \ }b\geq q_{k},

where QQ is any set of points in (Yk)ϵ(Y_{k})_{\epsilon} that meets all the connected components.

Proof.

For k=1k=1 the result follows by Proposition 69 because Y1Y_{1} is metrically conical. For k>1k>1 we consider the fibration

ξ|Yk:YkDϵ.\xi|_{Y_{k}}:Y_{k}\to D^{*}_{\epsilon}.

Restricting over the boundary Dϵ\partial D^{*}_{\epsilon} we have a fibration

ξ|(Yk)ϵ:(Yk)ϵDϵ\xi|_{(Y_{k})_{\epsilon}}:(Y_{k})_{\epsilon}\to\partial D^{*}_{\epsilon}

whose fibre is a surface FiF_{i}. The surface FiF_{i} decomposes in connected components as

Fi=j=1Nk=1MjFi,j,k,F_{i}=\coprod_{j=1}^{N}\coprod_{k=1}^{M_{j}}F_{i,j,k},

where the components Fi,j,1,,Fi,j,MjF_{i,j,1},...,F_{i,j,M_{j}} are interchanged cyclically by the monodromy. Two components Fi,j,kF_{i,j,k} and Fi,j,kF_{i,j^{\prime},k^{\prime}} are in the same connected component of YkY_{k} if and only if j=jj=j^{\prime}. In what follows we complete the proof for the special case N=1N=1 and M1=1M_{1}=1. The general case is exactly the same with some more notational complication, and this special case the Lipschitz geometry appears in a more transparent way.

Consider the decomposition

(22) Dϵ=Dϵ{(t,0),t>0}Dϵ{(t,0),t<0}D^{*}_{\epsilon}=D^{*}_{\epsilon}\setminus\{(t,0),t>0\}\cup D^{*}_{\epsilon}\setminus\{(t,0),t<0\}

and pullback this decomposition of DϵD^{*}_{\epsilon} by ξ\xi to a decomposition of YkY_{k}. Each of the two pieces of the decomposition is connected and admits the bb-cone over FiF_{i} as a metric deformation retract: this means that the bb-cone of FiF^{\prime}_{i} is included in each of this pieces and that the inclusion is a metric homotopy equivalence in the sense of Definition 41. The intersection of the two pieces splits as the disjoint union of two connected components, which are the preimages by ξ|Yk\xi|_{Y_{k}} of the two connected components of Dϵ{(t,0),t}D^{*}_{\epsilon}\setminus\{(t,0),t\in{\mathbb{R}}\}. Each of these connected components admits the bb-cone over FiF_{i} as a metric deformation retract.

Therefore, by Proposition 42 and Proposition 69 the bMDb-MD of each of the two pieces and of the connected components of the intersection between them of the decomposition is equal to the topological fundamental group of FiF_{i}.

We check hypothesis ()b(*)_{b} and ()b(**)_{b} in Theorem 55 as in the previous Proposition. Then, applying Theorem 55 to the decomposition and the Seifert van-Kampen Theorem for groupois to the corresponding decomposition of (Yk)ϵ(Y_{k})_{\epsilon}, one observe that both colimit computations are the same and the result is proven. ∎

Lemma 75.

There is an isomorphism of fundamental groupoids

MDπ1b(Z~i,j,𝔥(Q))π1((Z~i,j)ϵ,Q)for bqj>qi,MD\pi^{b}_{1}(\widetilde{Z}_{i,j},\mathfrak{h}(Q))\cong\pi_{1}((\widetilde{Z}_{i,j})_{\epsilon},Q)\ \ \text{for \ }b\geq q_{j}>q_{i},

where QQ is any set of points in (Z~i,j)ϵ(\widetilde{Z}_{i,j})_{\epsilon} that meets all the connected components.

Proof.

The boundary of (Z~i,j)ϵ(\widetilde{Z}_{i,j})_{\epsilon} is a disjoint union of tori classified in two kinds: those such that the piece of (Z~i,j)ϵ(\widetilde{Z}_{i,j})_{\epsilon} induced by them by the conical structure collapse at rate qiq_{i}, and those that the associated conical piece collapse at rate qjq_{j}. Recall that qj>qiq_{j}>q_{i}. Denote by (Ti)ϵ(T_{i})_{\epsilon} the union of tori of qiq_{i} type, and by TiT_{i} its associated conical piece. The inclusion TiZ~i,jT_{i}\hookrightarrow\widetilde{Z}_{i,j} is a metric homotopy invariance according with Definition 41. This reduces the problem to prove the isomorphism

MDπ1b(Ti)π1((Ti)ϵ),MD\pi^{b}_{1}(T_{i})\cong\pi_{1}((T_{i})_{\epsilon}),

but this is entirely analogous to the proof of the previous lemma. ∎

Exactly the same proof yields:

Lemma 76.

There are isomorphisms of fundamental groupoids

MDπ1b(Ck,𝔥(Q))π1((Ck)ϵ,Q),MD\pi^{b}_{1}(C_{k}^{-},\mathfrak{h}(Q^{-}))\cong\pi_{1}((C_{k}^{-})_{\epsilon},Q^{-}),
MDπ1b(Ck+,𝔥(Q+))π1((Ck+)ϵ,Q+)for bqkMD\pi^{b}_{1}(C_{k}^{+},\mathfrak{h}(Q^{+}))\cong\pi_{1}((C_{k}^{+})_{\epsilon},Q^{+})\ \ \text{for \ }b\geq q_{k}

where QQ^{*} is any set of points in (Ck)ϵ(C_{k}^{*})_{\epsilon} that meets all the connected components.

To study the bb-MD fundamental groupoid of the piece U~>b\widetilde{U}_{>b} we define the space (V~>b)ϵ(\widetilde{V}_{>b})_{\epsilon} as follows. Let us start considering the fibration

ξ|(U>b)ϵ:(U>b)ϵDϵ\xi|_{(U_{>b})_{\epsilon}}:(U_{>b})_{\epsilon}\to\partial D^{*}_{\epsilon}

with fibre F>bF_{>b}. Consider the decomposition in connected components

F>b=j=1Nk=1MjF>b,j,k,F_{>b}=\coprod_{j=1}^{N}\coprod_{k=1}^{M_{j}}F_{>b,j,k},

where the components F>b,j,1,,F>b,j,MjF_{>b,j,1},...,F_{>b,j,M_{j}} are interchanged cyclically by the monodromy. Two components F>b,j,kF_{>b,j,k} and F>b,j,kF_{>b,j^{\prime},k^{\prime}} are in the same connected component of (U>b)ϵ(U_{>b})_{\epsilon} if and only if j=jj=j^{\prime}. We denote the decomposition of (U>b)ϵ(U_{>b})_{\epsilon} in connected components by

(U>b)ϵ=j=1N(U>b)ϵj.(U_{>b})_{\epsilon}=\coprod_{j=1}^{N}(U_{>b})_{\epsilon}^{j}.

For each jj there exists a fibration

ηj:(U>b)ϵj𝕊1\eta_{j}:(U_{>b})_{\epsilon}^{j}\to{\mathbb{S}}^{1}

with connected fibre and a unique covering ρj:𝕊1Dϵ\rho_{j}:{\mathbb{S}}^{1}\to\partial D^{*}_{\epsilon} of degree MjM_{j} such that ρjηj=ξ|[(U>b)ϵ]j\rho_{j}{\circ}\eta_{j}=\xi|_{[(U_{>b})_{\epsilon}]_{j}}. Define

(V~>b)ϵ:=(U~>b)ϵ(U>b)ϵj=1NCyl(ηj)(\widetilde{V}_{>b})_{\epsilon}:=(\widetilde{U}_{>b})_{\epsilon}\cup_{(U_{>b})_{\epsilon}}\coprod_{j=1}^{N}Cyl(\eta_{j})

where Cyl(ηj)Cyl(\eta_{j}) is the mapping cylinder of and (U>b)ϵ\cup_{(U_{>b})_{\epsilon}} denotes the gluing of each piece Cyl(ηj)Cyl(\eta_{j}) along (U>b)ϵCyl(ηj)(U_{>b})_{\epsilon}\cap Cyl(\eta_{j}).

Lemma 77.

There is an isomorphism of fundamental groupoids

MDπ1b(U~>b,𝔥(Q))π1((V~>b)ϵ,(U~>b)ϵ)MD\pi^{b}_{1}(\widetilde{U}_{>b},\mathfrak{h}(Q))\cong\pi_{1}((\widetilde{V}_{>b})_{\epsilon},(\widetilde{U}_{>b})_{\epsilon})

where QQ is any set of points in (U~>b)ϵ(\widetilde{U}_{>b})_{\epsilon} that meets all the connected components.

Proof.

In order to compute the left hand side notice that the ξ|U~>b:U~>bDϵ\xi|_{\widetilde{U}_{>b}}:\widetilde{U}_{>b}\to D^{*}_{\epsilon} induces a decomposition of U~>b\widetilde{U}_{>b} by pullback of the decomposition (22). Arguments similar to the proof of Proposition 73 show that Seifert- van Kampen Theorem 55 can be applied to this decomposition.

Like in the proof of Lemma 74 we complete the proof for the special case N=1N=1 and M1=1M_{1}=1, since the general case is similar.

Then, as in the previous two lemmata, each of the two pieces of the decomposition, and each of the two connected component of the intersection between them is metrically homotopy equivalent to a connected component of the preimage B=(ξ|U~>b)1({(t,0):t>0}B=(\xi|_{\widetilde{U}_{>b}})^{-1}(\{(t,0):t>0\} of the ray {(t,0):t>0}\{(t,0):t>0\}. It is clear that BB is bb-contractible.

In order to compute the right hand side observe that, similarly, we have a fibration ξϵ:(V~>b)ϵDϵ\xi_{\epsilon}:(\widetilde{V}_{>b})_{\epsilon}\to\partial D^{*}_{\epsilon}, and the topological Seifert-van Kampen Theorem for groupoids can be applied to the decomposition of (V~>b)ϵ(\widetilde{V}_{>b})_{\epsilon} obtained by pullback of the restriction of decomposition  (22) to Dϵ\partial D^{*}_{\epsilon}. Each connected component of any finite intersection of subsets of the decomposition is contractible.

Comparing the colimit computations for the left and right hand sides we conclude. ∎

Finally, the computation of the whole MDπ1b(X)MD\pi^{b}_{1}(X) can be codified in the topology of the following space, which we can call it the (b,1)(b,1)-homotopy model of XX:

(23) Xϵb:=Xϵ(U>b)ϵCyl(ξ|(U>b)ϵ)=XϵU~>b(V~>b)ϵ.X^{b}_{\epsilon}:=X_{\epsilon}\cup_{(U_{>b})_{\epsilon}}Cyl(\xi|_{(U_{>b})_{\epsilon}})=X_{\epsilon}\cup_{\widetilde{U}_{>b}}(\widetilde{V}_{>b})_{\epsilon}.

The space XϵbX^{b}_{\epsilon} can be understood as the result of fibrewise identifying to a point the connected components of the part of the fibres of ξ\xi that collapse to a rate higher than bb. It has the homotopy type of a plumbed 33-manifold in which several circles are identified (a “branched 33-manifold” in the language of [3]).

Observe that if bbb\geq b^{\prime} we have a natural continuous map

αb,b:XϵbXϵb.\alpha_{b,b^{\prime}}:X^{b}_{\epsilon}\to X^{b^{\prime}}_{\epsilon}.
Theorem 78.

Let (X,0,dinn)(X,0,d_{inn}) be a normal surface singularity wuth the inner metric. Let PP be a set of points in XϵX_{\epsilon} that meets the interior of every connected components of YkY_{k}, Ck+C_{k}^{+}, CkC_{k}^{-} and every ZijZ_{ij}. Let 𝔓:=𝔥(P)\mathfrak{P}:=\mathfrak{h}(P) be defined after (21). Then π1(Xϵb,P)\pi_{1}(X^{b}_{\epsilon},P) is isomorphic to MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) for any b1b\geq 1.

If pp is any point in XϵX_{\epsilon} and 𝔭\mathfrak{p} any l.v.a. point in XX there is an isomorphism of 𝔹{\mathbb{B}}-groups from

π1(Xϵb,p)π1(Xϵb,p)...\to\pi_{1}(X^{b}_{\epsilon},p)\to\pi_{1}(X^{b^{\prime}}_{\epsilon},p)\to...

to MDπ1(X,0,dinn,𝔭)MD\pi_{1}^{\star}(X,0,d_{inn},\mathfrak{p}), that is to

MDπ1b(X,𝔭)MDπ1b(X,𝔭)....\to MD\pi_{1}^{b}(X,\mathfrak{p})\to MD\pi_{1}^{b^{\prime}}(X,\mathfrak{p})\to....
Proof.

The proof consists in comparing the computation of MDπ1b(X,𝔓)MD\pi_{1}^{b}(X,\mathfrak{P}) by the MDMD Seifert-van Kampen Theorem, Theorem 55 with the coverings 𝒰b{\mathcal{U}}^{b} in (19), and of π1(Xϵb,P)\pi_{1}(X^{b}_{\epsilon},P) by the topological Seifert van-Kampen Theorem associated with the covering obtained by restriction of (19), and use Lemmata 747576 and 77. ∎

Using the covering (19) and the models (23) we can also compute the MD homology of a complex surface singularity:

Theorem 79.

Let (X,0,dinn)(X,0,d_{inn}) be a normal surface singularity with the inner metric. For every nn\in{\mathbb{N}} we have isomorphisms from

Hn(Xϵb)H1(Xϵb)...\to H_{n}(X^{b}_{\epsilon})\to H_{1}(X^{b^{\prime}}_{\epsilon})\to...

to MDHn(X,0,dinn,𝔭)MDH_{n}^{\star}(X,0,d_{inn},\mathfrak{p}), that is to

MDHnb(X)MDH1b(X)....\to MDH_{n}^{b}(X)\to MDH_{1}^{b^{\prime}}(X)\to....
Proof.

The proof is similar to the previous Theorem, but we have to replace Seifert- van Kampen arguments by Mayer-Vietoris ones. We modify slightly the covering: consider

(24) 𝒰b:={Yk:qkb}{Z~i,j:qi<qjb}{U~>b.}{\mathcal{U}}^{*}_{b}:=\{{Y}^{*}_{k}:q_{k}\leq b\}\cup\{\widetilde{Z}^{*}_{i,j}:q_{i}<q_{j}\leq b\}\cup\{\widetilde{U}^{*}_{>b}.\}

defined as follows. Split the annuli CkC_{k}^{\bullet}, for {+,}\bullet\in\{+,-\} in 33 equal subannuli

Ck=Ck,YCk,middleCk,Z,C_{k}^{\bullet}=C_{k}^{\bullet,Y}\cup C_{k}^{\bullet,\mathrm{middle}}\cup C_{k}^{\bullet,Z},

where Ck,YC_{k}^{\bullet,Y} is the third of the annulus adjacent to the YY-piece and Ck,ZC_{k}^{\bullet,Z} is the piece of the annulus adjacent to the ZZ-piece. Define

Zij:=ZijCj,ZCj,middleCi+,ZCi+,middle,{Z}_{ij}^{*}:=Z_{ij}\cup C_{j}^{-,Z}\cup C_{j}^{-,\mathrm{middle}}\cup C_{i}^{+,Z}\cup C_{i}^{+,\mathrm{middle}},
Yk:=YkCj,YCj,middleCi+,YCi+,middle,{Y}^{*}_{k}:=Y_{k}\cup C_{j}^{-,Y}\cup C_{j}^{-,\mathrm{middle}}\cup C_{i}^{+,Y}\cup C_{i}^{+,\mathrm{middle}},

and U~>b\widetilde{U}_{>b}^{*} similarly.

This is a bb-covering as in Definition 92 in [8]: the subsets extending each of the subsets of the cover and their finite intersections are obtained adding the relevant 1/31/3-pieces of the corresponding annuli, and checking that such a choice works follows is a simple use of the MD Homology invariance by metric homotopy. Then, we can apply repeately the Mayer-Vietoris type theorem, Theorem 98 in [8] in order to compute the bMDb-MD Homology of XX. Computing the Homology groups of XϵbX_{\epsilon}^{b} applying repeatedly the Mayer-Viertoris sequence for ordinary homology for the decomposition of XϵbX_{\epsilon}^{b} corresponding to the decomposition of XX defined above, and comparing the computations yields the result. ∎

Let us finish with a few open problems:

Problem 80.

Let (X,O)(X,O) be a normal surface singularity,

  1. (1)

    If (X,O)(X,O) is not a cyclic quotient, does the MD-fundamental group determine the inner geometry of a normal complex surface singularity? This is motivated by the corresponding statement, due to Waldhausen, for the topology.

  2. (2)

    Find a homotopy model computing MDπ1b(X,O,dout)MD\pi_{1}^{b}(X,O,d_{out}).

  3. (3)

    If the natural homomorphism MDπ1b(X,O,dinn)MDπ1b(X,O,dout)MD\pi_{1}^{b}(X,O,d_{inn})\to MD\pi_{1}^{b}(X,O,d_{out}) is an isomorphism for every bb, is (X,O)(X,O) Lipschitz normally embedded?

Problem 81.

Compute MDπ1bMD\pi_{1}^{b} and MDHbMDH_{\star}^{b} for any Brieskorn-Pham singularity. The higher homotopy group computation should be very hard, since it contains the homotopy groups of spheres.

References

  • [1] Bernig, A. and Lytchak, A. Tangent spaces and Gromov-Hausdorff limits of subanalytic spaces. (English summary) J. Reine Angew. Math. 608 (2007), 115.
  • [2] Birbrair, L. , Mostowski, T., Normal embeddings of semialgebraic sets. Michigan Math. J. 47 (2000), 125132.
  • [3] Birbrair L., Neumann W., Pichon A., The thick-thin decomposition and the bilipschitz classification of normal surface singularities. Acta Math., 212 (2014), 199-256.
  • [4] Brown, R. and Razak, A. ‘A van Kampen theorem for unions of non-connected spaces’, Archiv. Math. 42 (1984) 85-88
  • [5] Coste, M. An Introduction to O-minimal Geometry. Institut de Recherche Mathématique de Rennes (1999). https://perso.univ-rennes1.fr/michel.coste/polyens/OMIN.pdf.
  • [6] van den Dries, L. A generalization of the Tarski-Seidenberg theorem, and some nondefinability results Bull. Amer. Math. Soc. (N.S.), vol. 15 (1986), n. 2, 189-193.
  • [7] van den Dries, L. Tame Topology and O-minimal Structures. London Mathematical Society Lecture Note Series, vol. 248, Cambridge University Press (1998).
  • [8] Fernández de Bobadilla, J. , Heinze, S. , Pe Pereira, M. Moderately Discontinuous Homology
  • [9] Hatcher, A. Algebraic Topology Cambridge University Press, 2002.
  • [10] Heinze, S. Moderately discontinuous algebraic topology for metric subanalytic germs PhD Thesis, UCM 2019. https://eprints.ucm.es/59485/1/T41827.pdf.