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

Semi-coarse Spaces: Fundamental Groupoid and the van Kampen Theorem

Jonathan Treviño-Marroquín
Abstract.

In algebraic topology, the fundamental groupoid is a classical homotopy invariant which is defined using continuous maps from the closed interval to a topological space. In this paper, we construct a semi-coarse version of this invariant, using as paths a finite sequences of maps from 1\mathbb{Z}_{1} to a semi-coarse space, connecting their tails through semi-coarse homotopy. In contrast to semi-coarse homotopy groups, this groupoid is not necessarily trivial for coarse spaces, and, unlike coarse homotopy, it is well-defined for general semi-coarse spaces. In addition, we show that the semi-coarse fundamental groupoid which we introduce admits a version of the Van Kampen Theorem.

This work was supported by the grant N62909-19-1-2134 from the US Office of Naval Research Global and the Southern Office of Aerospace Research and Development of the US Air Force Office of Scientific Research. The author was also supported by the CONACYT postgraduate studies scholarship number 839062, and this document is part of his PhD project.

1. Introduction

The category of semi-coarse spaces is one of several topological constructs that has been taken up to study algebraic-topological invariants on graphs and semipseudometric spaces, which together serve as a foundation for topological data analysis. Other categories which have been proposed for this purpose include Čech closure spaces [Bubenik_Milicevic_2021b, Rieser_2021] and Choquet Spaces [Rieser_arXiv_2022]. Semi-coarse spaces generalize Roe’s axiomatization of coarse spaces [Roe_2003] by removing the product of sets axiom. This is fundamental because it provides us with a collection of interval objects for semi-coarse spaces, and we can define homotopy maps in this category [rieser2023semicoarse]. Additionally, unlike in coarse geometry, compact (metric) spaces admit semi-coarse structures which are no longer necessarily equivalent to the point, allowing one to coarsen a compact metric space up to a pre-defined scale [Zava_2019] without making it trivial as a semi-coarse space. Canonical examples of semi-coarse spaces are the (undirected) graphs, such as the cyclic graphs and the Cayley graph of the integers with the unit as generator (1\mathbb{Z}_{1}), and semipseudometric spaces, in particular those constructed from metric spaces with a privileged scale. The loss of the product axiom for coarse spaces produces two possible definitions of bounded sets and fails to fulfill that the union of bounded sets is bounded in connected spaces. Hence, we use bornological maps as our morphisms, which avoids the use of bounded sets to define, but complicates the adaptation of coarse homotopy to semi-coarse spaces (for more on this, see Appendix A).

In previous work, joint with Antonio Rieser [rieser2023semicoarse], we defined homotopy groups for arbitrary semi-coarse spaces. There we observe that, when applied to coarse spaces, π0\pi_{0} quantifies the number of (coarsely) connected components, and the rest of the homotopy groups vanishe. In this paper, we construct an invariant which may be non-trivial on both coarse and non-coarse semi-coarse spaces, the semi-coarse fundamental groupoid, inspired by the one in topology [Brown_Higgins_Sivera_2011] and by the visual boundary in semi-geodesic spaces [Kitzmiller_2009]. This construction generalizes the semi-coarse fundamental group, and reduces to it under some assumptions. While the definition of the new fundamental groupoid is slightly technical, it is indeed a semi-coarse invariant which is non-trivial on coarse spaces and is able to distinguish them. Moreover, under some specific cases, we have an analogue to van Kampen theorem for semi-coarse spaces. In related work, the fundamental groupoid has been also realized for graphs in AA-theory [kapulkin2023fundamental] and for ×\times-homotopy [Chih_Scull_2022].

The article is organized as follows. We introduce the category of semi-coarse spaces in Section 2. There, we recall some results and definitions of [rieser2023semicoarse], and we introduce semi-coarse homotopy and the fundamental group. In semi-coarse spaces, we do not have a canonical collection of sets with which to construct useful covers, i.e. like open sets in topological spaces, and in Section 3 we partially resolve this problem by defining when a pair of subsets well-split a space (3.6) and we establish several relations between well-splitting, coarse spaces, and connectedness. The definition of well-splitting provides us with a sufficient condition for proving a van Kampen theorem for the semi-coarse fundamental groupoid. In Section 4, we define strings of paths and an equivalence relation on them, and we use these to construct the semi-coarse fundamental groupoid.

In Section 5, we begin with the main results of the paper, and define the relative fundamental groupoid and prove a semi-coarse version of the Van Kampen Theorem give some mild hypotheses. Finally, in Appendix A we discuss several issues which arise when trying to generalize coarse homotopy to both coarse and semi-coarse spaces. We do not discard the possbility that this notion can be extended, but it appears that such a generalization is difficult to construct.

2. Semi-coarse Spaces and its Homotopy

Our first first approximation was looking for a space where we can take some ideas of coarse geometry to study any (undirected) graphs and metric spaces with privileged scale. This category received the name of semi-coarse spaces [Zava_2019] and we developed the homotopy with some ideas of [Babson_etal_2006] using the cartesian product instead the inductive (or box) product. This section is a summary of [rieser2023semicoarse], we enunciate the definitions and results we use in the subsequent sections.

Definition 2.1 (Semi-coarse space; [rieser2023semicoarse]).

Let XX be a set, and let 𝒱𝒫(X×X)\mathcal{V}\subset\mathcal{P}(X\times X) be a collection of subsets of X×XX\times X which satisfies

  1. (sc1)

    ΔX𝒱\Delta_{X}\in\mathcal{V},

  2. (sc2)

    If B𝒱B\in\mathcal{V} and ABA\subset B, then A𝒱A\in\mathcal{V},

  3. (sc3)

    If A,B𝒱A,B\in\mathcal{V}, then AB𝒱A\cup B\in\mathcal{V},

  4. (sc4)

    If A𝒱A\in\mathcal{V}, then A1𝒱A^{-1}\in\mathcal{V}.

We call 𝒱\mathcal{V} a semi-coarse structure on XX, and we say that the pair (X,𝒱)(X,\mathcal{V}) is a semi-coarse space. If, in addition, 𝒱\mathcal{V} satisfies

  1. (sc5)

    If A,B𝒱A,B\in\mathcal{V}, then AB𝒱A\circ B\in\mathcal{V}.

then 𝒱\mathcal{V} will be called a coarse structure, and (X,𝒱)(X,\mathcal{V}) will be called a coarse space, as in [Roe_2003].

A map f:XYf:X\rightarrow Y between the coarse spaces (X,𝒱)(X,\mathcal{V}) and (Y,𝒲)(Y,\mathcal{W}) is called (𝒱,𝒲)(\mathcal{V},\mathcal{W})-bornologous if f×f(V){(f(a),f(b))(a,b)V}𝒲f\times f(V)\coloneqq\{(f(a),f(b))\mid(a,b)\in V\}\in\mathcal{W} for every V𝒱V\in\mathcal{V}. We observe that semi-coarse spaces with bornologous maps is a category.

Semi-coarse spaces is a topological construct, the practical effect for this paper is that final and initial structures are defined for every collection of semi-coarse spaces [Zava_2019] (to see the formal definition of Topological Constructs and some properties see [Beattie_Butzmann_2002, Preuss_2002]). Some important structures are: subspace, product space, quotients and disjoint union space.

Definition 2.2 (Semi-coarse Subspace; 2.2.3 [rieser2023semicoarse]).

Let (X,𝒱)(X,\mathcal{V}) be a semi-coarse space and let YXY\subset X. (Y,𝒱Y)(Y,\mathcal{V}_{Y}) is called a semi-coarse subdspace of XX with 𝒱Y\mathcal{V}_{Y} the collection

{V(Y×Y)V𝒱}\displaystyle\{V\cap(Y\times Y)\mid V\in\mathcal{V}\}

In the following sense, bornologous maps might be seen locally: If all of the controlled sets might be as the union of the intersections with sets of the way Xi×XiX_{i}\times X_{i}, then we can observe if the maps are bornologous there, and conclude they are bornologous in the whole.

Proposition 2.3 ([rieser2023semicoarse] 2.2.4).

Let (X,𝒱)(X,\mathcal{V}) and (Y,𝒲)(Y,\mathcal{W}) be semi-coarse spaces, and suppose that (Xi,𝒱i)(X,𝒱)(X_{i},\mathcal{V}_{i})\subset(X,\mathcal{V}), i{1,,n}i\in\{1,\ldots,n\}, are subspaces of (X,𝒱)(X,\mathcal{V}) such that i=1nXi=X\cup_{i=1}^{n}X_{i}=X and every set V𝒱V\in\mathcal{V} may be written in the form

(1) V=i=1nVi\displaystyle V=\bigcup_{i=1}^{n}V_{i}

where each Vi𝒱iV_{i}\in\mathcal{V}_{i}. Now suppose that f:XYf:X\rightarrow Y is a map such that the restrictions fXi:(Xi,𝒱i)(Y,𝒲)f\mid_{X_{i}}:(X_{i},\mathcal{V}_{i})\rightarrow(Y,\mathcal{W}) are bornologous for all i{1,,n}i\in\{1,\ldots,n\}. Then f:(X,𝒱)(Y,𝒲)f:(X,\mathcal{V})\rightarrow(Y,\mathcal{W}).

Definition 2.4.

[Semi-coarse Product; 2.3.7 [rieser2023semicoarse]] Let (X,𝒱)(X,\mathcal{V}) and (Y,𝒲)(Y,\mathcal{W}) be semi-coarse spaces. (X×Y,𝒱×𝒲)(X\times Y,\mathcal{V}\times\mathcal{W}) is called the product space, and 𝒱×𝒲\mathcal{V}\times\mathcal{W} the product structure, where U𝒱×𝒲U\in\mathcal{V}\times\mathcal{W} if there exists V𝒱V\in\mathcal{V} and W𝒲W\in\mathcal{W} such that UVWU\subset V\boxtimes W with

VW{(v,w,v,w)(v,v)V,(w,w)W}\displaystyle V\boxtimes W\coloneqq\{(v,w,v^{\prime},w^{\prime})\mid(v,v^{\prime})\in V,(w,w^{\prime})\in W\}

Since semi-coarse spaces are a topological construct, inductive product is also defined and it defines another kind of homotopy (see examples in other categories in [Babson_etal_2006, Bubenik_Milicevic_2021b].) For the objectives in this paper, it is enough to work with (categorical) product space; it could be interesting to explore the later definitions with that product.

Definition 2.5 (Quotient Space; 2.4.1 and 2.4.2 [rieser2023semicoarse]).

Let (X,𝒱)(X,\mathcal{V}) be a semi-coarse space, let YY be a set, and let g:XYg:X\rightarrow Y be a surjective function. We define

𝒱g{(g×g)(V)V𝒱}.\displaystyle\mathcal{V}_{g}\coloneqq\{(g\times g)(V)\mid V\in\mathcal{V}\}.

Then (Y,𝒱g)(Y,\mathcal{V}_{g}) is called the quotient space (or semi-coarse space inductively generated by the function gg) and 𝒱g\mathcal{V}_{g} is called the quotient structure (or semi-coarse structure inductively generated by the function gg).

Theorem 2.6 ([rieser2023semicoarse] 2.4.3).

Let (Y,𝒱g)(Y,\mathcal{V}_{g}) be a quotien semi-coarse space inductively generated by the function g:XYg:X\rightarrow Y and let (Z,𝒵)(Z,\mathcal{Z}) be a semi-coarse space. A function f:(Y,𝒱g)(Z,𝒵)f:(Y,\mathcal{V}_{g})\rightarrow(Z,\mathcal{Z}) is bornologous if and only if fg:XZf\circ g:X\rightarrow Z is bornologous.

Definition 2.7 (Disjoint Union; 2.5.1 and 2.5.2 [rieser2023semicoarse]).

Let {(Xλ,𝒱λ)}λΛ\{(X_{\lambda},\mathcal{V}_{\lambda})\}_{\lambda\in\Lambda} be a collection of semi-coarse spaces indexed by the set Λ\Lambda, and let λΛ𝒱λ\sqcup_{\lambda\in\Lambda}\mathcal{V}_{\lambda} be the collection of sets of the form

λΛAλ,\displaystyle\bigsqcup_{\lambda\in\Lambda^{\prime}}A_{\lambda},

where |Λ|<|\Lambda^{\prime}|<\infty and each Aλ𝒱λA_{\lambda}\in\mathcal{V}_{\lambda}. (Note that any given AλA_{\lambda} may be the empty set.)

We call λΛ𝒱λ\sqcup_{\lambda\in\Lambda}\mathcal{V}_{\lambda} the disjoint union semi-coarse structure, and (λΛXλ,λΛ𝒱λ)(\sqcup_{\lambda\in\Lambda}X_{\lambda},\sqcup_{\lambda\in\Lambda}\mathcal{V}_{\lambda}) the disjoint union of the semi-coarse spaces {(Xλ,𝒱λ)}λΛ\{(X_{\lambda},\mathcal{V}_{\lambda})\}_{\lambda\in\Lambda}.

Proposition 2.8.

Let {(Xλ,𝒱λ}\{(X_{\lambda},\mathcal{V}_{\lambda}\} be a collection of semi-coarse spaces indexed bu the set Λ\Lambda and (Z,𝒵)(Z,\mathcal{Z}) be a semi-coarse space. Define iλ:XλXi_{\lambda}:X_{\lambda}\rightarrow X such that iλ(x)=(x,λ)i_{\lambda}(x)=(x,\lambda). Then g:(Xλ,𝒱λ)(Z,𝒵)g:(\sqcup X_{\lambda},\sqcup\mathcal{V}_{\lambda})\rightarrow(Z,\mathcal{Z}) is a bornologous maps if and only if giλ:XλZg\circ i_{\lambda}:X_{\lambda}\rightarrow Z is bornologous for every λΛ\lambda\in\Lambda.

Proof.

By defintion, iλi_{\lambda} sends controlled sets in controlled set, then it is a bornologous map. Thus, if gg is bornologous, we have that giλg\circ i_{\lambda} is bornologous for every λΛ\lambda\in\Lambda.

On the other hand, consider that giλg\circ i_{\lambda} is bornologous for every λ\lambda. Take a controlled set AA of X\sqcup X, then there exists λ1,,λnΛ\lambda_{1},\ldots,\lambda_{n}\in\Lambda and Aλj𝒱iA_{\lambda_{j}}\in\mathcal{V}_{i} such that A=AλjA=\sqcup A_{\lambda_{j}}. Since giλig\circ i_{\lambda_{i}} is bornologous, then g(Aλj)𝒵g(A_{\lambda_{j}})\in\mathcal{Z}, then g(Aλj)𝒵\bigcup g(A_{\lambda_{j}})\in\mathcal{Z}. Therefore gg is bornologous. ∎

Through the disjoint union, we can generate the coarser coarse space which contains a semi-coarse space.

Definition 2.9 (Product Extension; 2.5.9 and 2.5.10 [rieser2023semicoarse]).

Let (X,𝒱)(X,\mathcal{V}) be a semi-coarse space, and define

𝒱PE{CX×X:A,B𝒱 with CAB}.\displaystyle\mathcal{V}^{PE}\coloneqq\{C\subset X\times X:\exists A,B\in\mathcal{V}\text{ with }C\subset A\circ B\}.

We call the structure 𝒱EP\mathcal{V}^{EP} the set product extension of 𝒱\mathcal{V}, and the ordered pair (X,𝒱EP)(X,\mathcal{V}^{EP}) is called the set product extension of (X,𝒱)(X,\mathcal{V}). For any kk\in\mathbb{N}, we recursively define 𝒱kPE\mathcal{V}^{kPE} to be the set product extension of 𝒱(k1)PE\mathcal{V}^{(k-1)PE}.

Observe that 𝒱𝒱PE\mathcal{V}\subset\mathcal{V}^{PE}.

Definition 2.10 (Product Extension; 2.5.14 and 2.5.15 [rieser2023semicoarse]).

Let (X,𝒱)(X,\mathcal{V}) be a semi-coarse space, and let {(X,𝒱kPE,ιij,}\{(X,\mathcal{V}^{kPE},\iota_{i}^{j},\mathbb{N}\} be the directed system of semi-coarse spaces such that all the ι:(X,𝒱iPE)(X,𝒱jPE)\iota:(X,\mathcal{V}^{iPE})\rightarrow(X,\mathcal{V}^{jPE}) are the identity map. We call that coarse structure the coarse structure induced by 𝒱\mathcal{V}, and which we denote by 𝒱\mathcal{V}^{\infty}.

To close this section, we introduce briefly the semi-coarse homotopy and the fundamental semi-coarse group. The first concept is necessary to define what a string is (4.5); the second one is to illustrate that all of the fundamental semi-coarse groups of XX are contained in the fundamental groupoid.

From here on out, we denote by 1\mathbb{Z}_{1} the semi-coarse space on \mathbb{Z} such that the controlled sets are subsets of {(i,i+j)i,j{1,0,1}}\{(i,i+j)\mid i\in\mathbb{Z},j\in\{-1,0,1\}\}.

Definition 2.11 (Semi-coarse Homotopy; 3.1.3 [rieser2023semicoarse]).

Let XX and YY semi-coarse spaces. f,g:XYf,g:X\rightarrow Y bornologous maps are homotopic if there exists H:X×1YH:X\times\mathbb{Z}_{1}\rightarrow Y bornologous and M,NM,N\in\mathbb{Z} such that H(x,k)=f(x)H(x,k)=f(x) for every kMk\leq M and H(x,k)=g(x)H(x,k)=g(x) for every kNk\geq N.

Definition 2.12 (Homotopy of Maps of Pairs and Triples; 3.1.6 [rieser2023semicoarse]).

Let XX and YY be semi-coarse spaces, let BAXB\subset A\subset X and DCYD\subset C\subset Y be endowed with the subspace structure, and let f,g:(X,A,B)(Y,C,D)f,g:(X,A,B)\rightarrow(Y,C,D) be bornologous maps of a triple, i.e. such that fACf\mid_{A}\subset C and fBDf\mid_{B}\subset D. We say that ff is relatively homotopic to gg and weite fscgf\simeq_{sc}g if and only if there is a homotopy H:X×1YH:X\times\mathbb{Z}_{1}\rightarrow Y such that HA×CH\mid_{A\times\mathbb{Z}}\subset C and HB×DH\mid_{B\times\mathbb{Z}}\subset D.

We define a homotopy between maps of pairs f,g:(X,A)(Y,C)f,g:(X,A)\rightarrow(Y,C) to be the homotopy between maps of a triple as above with B=AB=A and C=DC=D.

We summarize the building of the fundamental group in [rieser2023semicoarse] from section 3.2. Fix a semi-coarse space XX and consider all of the {0,,n}\{0,\ldots,n\} as subspace of 1\mathbb{Z}_{1}, and consider f:{0,,n}Xf:\{0,\ldots,n\}\rightarrow X as bornologous maps such that f(0)=f(n)=xf(0)=f(n)=x for a fixed xXx\in X. For every n<m0n<m\in\mathbb{N}_{0} we might extend the bornologous maps f:{0,,n}Xf:\{0,\ldots,n\}\rightarrow X to f:{0,,m}Xf^{\prime}:\{0,\ldots,m\}\rightarrow X such that f(i)=f(i)f^{\prime}(i)=f(i) for every 0in0\leq i\leq n and f(i)=f(n)f^{\prime}(i)=f(n) for every n<imn<i\leq m.

Calling that extension inmi_{n}^{m}. That construction produces the directed set

([f:{0,,n}X],inm,)\displaystyle([f:\{0,\ldots,n\}\rightarrow X],i_{n}^{m},\mathbb{N})

where we apply the direct limit which we call π1sc(X,x)\pi_{1}^{sc}(X,x). We name the classes there as [f][f] and we define the homotopy as [f]sc[g][f]\simeq_{sc}[g] if there are ff.

Finally, we define a product among these function. Let m,mm,m^{\prime}\in\mathbb{N} and suppose that f:{0,,m}Xf:\{0,\ldots,m\}\rightarrow X and g:{0,,m}Xg:\{0,\ldots,m^{\prime}\}\rightarrow X. We define the \star-product fg:Im+mXf\star g:I_{m+m^{\prime}}\rightarrow X such that

fg(j){f(j) if 0jmf(jm) if j>m.\displaystyle f\star g(j)\coloneqq\left\{\begin{array}[]{ll}f(j)&\text{ if }0\leq j\leq m\\ f(j-m)&\text{ if }j>m.\end{array}\right.

This map induces a product in π1sc(X,x)\pi_{1}^{sc}(X,x) and for Theorem 3.2.18 in [rieser2023semicoarse] we have that that collection is a groupo with \star-operation.

3. Connectedness and Splitting-well

Topological spaces have a considerable advantage over other categories: open sets. When we have an open cover {U,V}\{U,V\}, we already get that the pushout of UUVVU\leftarrow U\cap V\rightarrow V is exactly XX up isomorphisms; furthermore, we would know that XX is disconnected if UV=U\cap V=\varnothing. In general, semi-coarse spaces don not have a family of sets with such characteristics. To remedy this deficiency, we study the property that two subsets A,BA,B well-split XX (3.6). This kind of division help is an important tool to prove Van Kampen theorem for specific subsets.

Definition 3.1.

Let XX be a semi-coarse space. We say that XX is connected if for every x,yXx,y\in X there exists a path γ:{0,1,,n}X\gamma:\{0,1,\ldots,n\}\rightarrow X, {0,1,,n}\{0,1,\ldots,n\} subspace of 1\mathbb{Z}_{1}, such that γ(0)=x\gamma(0)=x and γ(n)=y\gamma(n)=y.

Definition 3.2.

Let A,B,CA,B,C be semi-coarse spaces and f:CAf:C\rightarrow A and g:CBg:C\rightarrow B be bornologous maps. We define PB/f(c)g(c)P_{B}/f(c)\sim g(c) is called the pushout of ACBAB/f(c)g(c)A\sqcup_{C}B\coloneqq A\sqcup B/f(c)\sim g(c) is called the pushout of A\textstyle{A}C\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces C\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}g\scriptstyle{g}B\textstyle{B}.

Proposition 3.3 (Universal Property of Pushout).

Let A,B,CA,B,C be semi-coarse spaces and f:CAf:C\rightarrow A and g:CBg:C\rightarrow B be bornologous maps. The diagram

is commutative, and for every other commutative diagram

we have that there exists a unique bornologous map h:ACBXh:A\sqcup_{C}B\rightarrow X such that j1=hi1j_{1}=hi_{1} and j2=hi2j_{2}=hi_{2}.

Proof.

Using the classical notation, AB=(A×{1})(B×{2})A\sqcup B=(A\times\{1\})\cup(B\times\{2\}). Moreover, we call i1(a)=[a,1]i_{1}(a)=[a,1] and i2(b)=[b,1]i_{2}(b)=[b,1].

We define the maps

h:ABX\displaystyle h:A\sqcup B\rightarrow X such that h(a,1)=j1(a) and h(b,2)=j2,\displaystyle\text{ such that }h(a,1)=j_{1}(a)\text{ and }h(b,2)=j_{2},
h¯:ACBX\displaystyle\overline{h}:A\sqcup_{C}B\rightarrow X such that h[a,1]=j1(a) and h[b,2]=j2.\displaystyle\text{ such that }h[a,1]=j_{1}(a)\text{ and }h[b,2]=j_{2}.

We want to prove that h¯\overline{h} is well-defined bornologous map, and it is the unique bornologous map such that j2=h¯i2j_{2}=\overline{h}\circ i_{2} and j1=h¯i1j_{1}=\overline{h}\circ i_{1}.

To prove that h¯\overline{h} is well-defined, we obtain the following cases:

  • If (a,1)[b,2](a,1)\in[b,2], then there exists cCc\in C such that f(c)=af(c)=a and g(c)=bg(c)=b. Thus h¯[a,1]=h¯i1(a)=h¯i1f(c)=h¯i2g(c)=h¯i2(b)=h¯[b,2]\overline{h}[a,1]=\overline{h}i_{1}(a)=\overline{h}i_{1}f(c)=\overline{h}i_{2}g(c)=\overline{h}i_{2}(b)=\overline{h}[b,2].

  • If (a,1)[a,1](a^{\prime},1)\in[a,1], then there exists c,cCc,c^{\prime}\in C such that f(c)=af(c)=a, f(c)=af(c^{\prime})=a^{\prime} and g(c)=g(c)g(c)=g(c^{\prime}). Thus h¯[a,1]=h¯[g(c),2]=h¯[g(c),3]=h¯[a,1]\overline{h}[a,1]=\overline{h}[g(c),2]=\overline{h}[g(c^{\prime}),3]=\overline{h}[a^{\prime},1].

We also have analogous cases for (b,2)[a,1](b,2)\in[a,1] and (b,2)[b,2](b^{\prime},2)\in[b,2], the procedure is completely analogous.

Now, we can prove that h¯\overline{h} is bornologous using the universal properties of the disjoint union and quotient, and observing that the following diagrams are commutative

Lastly, we observe that h¯\overline{h} is the unique map which satisfies that j2=h¯i2j_{2}=\overline{h}\circ i_{2} and j1=h¯i1j_{1}=\overline{h}\circ i_{1}. Suppose that there exists h:ACBXh^{\prime}:A\sqcup_{C}B\rightarrow X such that h[a,1]h¯[a,1]h^{\prime}[a,1]\neq\overline{h}[a,1], for some aAa\in A, then hi1(a)h¯i1(a)=j2(a)h^{\prime}i_{1}(a)\neq\overline{h}i_{1}(a)=j_{2}(a). Analogously if we take [b,2][b,2] with bBb\in B. ∎

Proposition 3.4.

Let (X,𝒱)(X,\mathcal{V}) a semi-coarse space. If {(x1,x2)}𝒱A𝒱AB𝒱B\{(x_{1},x_{2})\}\in\mathcal{V}_{A}\sqcup_{\mathcal{V}_{A\cap B}}\mathcal{V}_{B} and:

  1. (1)

    x1Ax_{1}\in A, x2Ax_{2}\notin A, then x1Bx_{1}\in B.

  2. (2)

    x1Bx_{1}\in B, x2Bx_{2}\notin B, then x1Ax_{1}\in A.

Proof.

Let (X,𝒱)(X,\mathcal{V}) a semi-coarse space, A,BXA,B\subset X which well-split XX and {(x1,x2)}𝒱A𝒱AB𝒱B\{(x_{1},x_{2})\}\in\mathcal{V}_{A}\sqcup_{\mathcal{V}_{A\cap B}}\mathcal{V}_{B}. Both cases are completely analogous; thus we prove just the first case. Therefore, consider that x1Ax_{1}\in A and x2Ax_{2}\notin A.

We call (Y,𝒲)(Y,\mathcal{W}) the subspace {1,2,3}\{1,2,3\} of 1\mathbb{Z}_{1}, and define the maps

j1:AY\displaystyle j_{1}:A\rightarrow Y such that j1(AB)=2 and j1(AB)=1\displaystyle\text{ such that }j_{1}(A\cap B)=2\text{ and }j_{1}(A-B)=1
j2:BY\displaystyle j_{2}:B\rightarrow Y such that j2(AB)=2 and j2(BA)=3\displaystyle\text{ such that }j_{2}(A\cap B)=2\text{ and }j_{2}(B-A)=3

Since {1,2}×{1,2},{2,3}×{2,3}𝒲\{1,2\}\times\{1,2\},\{2,3\}\times\{2,3\}\in\mathcal{W}, then j1,j2j_{1},j_{2} are bornologous maps. Moreover, the diagram

is commutative. Thus, since AABBA\sqcup_{A\cap B}B is a pushout, there exists a unique h:AABBYh:A\sqcup_{A\cap B}B\rightarrow Y bornologous such that j2=hi2j_{2}=hi_{2} and j1=hi1j_{1}=hi_{1}. Since hh is bornologus, then {h(x1),h(x2)}𝒲\{h(x_{1}),h(x_{2})\}\in\mathcal{W}. In addition, x1Ax_{1}\in A and x2BAx_{2}\in B-A, thus h(x2)=3h(x_{2})=3 and h(x1){1,2}h(x_{1})\in\{1,2\}. h(x1)1h(x_{1})\neq 1 because {(1,3)}𝒲\{(1,3)\}\notin\mathcal{W}. Therefore h(x1)=2h(x_{1})=2 and x1ABx_{1}\in A\cap B. ∎

Corollary 3.5.

Let XX be a semi-coarse space and A,BXA,B\subset X. Then every path γ\gamma in AABBA\sqcup_{A\cap B}B satisfies that

  • if γ(i)A\gamma(i)\in A and γ(i+1)A\gamma(i+1)\notin A, then γ(i)B\gamma(i)\in B, and

  • if γ(i)B\gamma(i)\in B and γ(i+1)B\gamma(i+1)\notin B, then γ(i)A\gamma(i)\in A.

Definition 3.6.

Let (X,𝒱)(X,\mathcal{V}) be a semi-coarse space and A,BXA,B\subset X non-empty. Then A,BA,B well-split X=ABX=A\cup B if for every x,y,yXx,y,y^{\prime}\in X such {(x,y),(x,y)}𝒱AABB\{(x,y),(x,y^{\prime})\}\in\mathcal{V}-A\sqcup_{A\cap B}B and {(y,y)}AABB\{(y,y^{\prime})\}\in A\sqcup_{A\cap B}B:

  1. (1)

    There exists a path γ,γ:{0,1,2}X\gamma,\gamma^{\prime}:\{0,1,2\}\rightarrow X in AABBA\sqcup_{A\cap B}B such that γ(0)=γ(0)=x\gamma(0)=\gamma^{\prime}(0)=x, γ(2)=y\gamma(2)=y, γ(2)=y\gamma^{\prime}(2)=y^{\prime} and γ(1)=γ(1)\gamma(1)=\gamma^{\prime}(1).

  2. (2)

    The set {γ(1)γ is a path in AABB such thatγ(0)=x,γ(2)=y}\{\gamma(1)\mid\gamma\text{ is a path in }A\sqcup_{A\cap B}B\text{ such that}\gamma(0)=x,\gamma(2)=y\} is connected as subspace of XX.

there exists a path γ:{0,1,2}X\gamma:\{0,1,2\}\rightarrow X in AABBA\sqcup_{A\cap B}B such that λscγ\lambda\simeq_{sc}\gamma.

xxxx^{\prime}yyyy^{\prime}wwww^{\prime}
Figure 1. Well-splitting example: Red points are ABA-B, blue points are BAB-A and violet points are ABA\cap B. XX with the semi-coarse space induce by the graph.

Condition two is equivalent to say that every f:{0,1,2}Xf:\{0,1,2\}\rightarrow X with f(0)=xf(0)=x and f(2)=yf(2)=y are homotopic. This implies that every A,BA,B with the first condition in a coarse space well-split the space. Indeed, the reason to include these both conditions is to include coarse spaces in our analyze.

Proposition 3.7.

Let XX be a semi-coarse space and A,BCA,B\subset C. If XX is (semi-coarse) homeomorphic to AABBA\sqcup_{A\cap B}B, then A,BXA,B\subset X well-split XX.

Example in Figure 1 is a clearly example that the other direction is not necessary true. It might be tempting to define take only A,BXA,B\subset X such that XX and XX and AABBA\sqcup_{A\cap B}B are semi-coarse isomorphic; however, we might leave out the interesting coarse cases.

Consider for example \mathbb{R} with its coarse structure induced by the metric, and A[0,)A\coloneqq[0,\infty), B(,0]B\coloneqq(-\infty,0]. AB=XA\cup B=X, but AABBA\sqcup_{A\cap B}B is not a coarse space and thus is not isomorphic to XX (see Figure 2 to compare both spaces).

\mathbb{R}AABBA\sqcup_{A\cap B}B\mathbb{R}\mathbb{R}XX\mathbb{R}
Figure 2. Every controlled set in AABBA\sqcup_{A\cap B}B and XX, respectively, is contained in some red area. Observe that there are not controlled sets in the second and fourth quadrants in AABBA\sqcup_{A\cap B}B, in contrast with XX.
Proposition 3.8.

Let (X,𝒱)(X,\mathcal{V}) be a semi-coarse space and A,BXA,B\subset X non-empty. XX is disconnected if and only if there exist A,BA,B which well-split XX such that AB=A\cap B=\varnothing.

Proof.

Let (X,𝒱)(X,\mathcal{V}) be a semi-coarse space and A,BXA,B\subset X non-empty.

()(\Rightarrow) Suppose that XX is disconnected and xXx\in X. Take AA as the component of xx and BXAB\coloneqq X-A. Observe that AABB=AB=XA\sqcup_{A\cap B}B=A\sqcup B=X as semi-coarse spaces. Thus, A,BA,B well-split XX.

()(\Leftarrow) Suppose that A,BA,B well-split XX and AB=A\cap B=\varnothing. Let aAa\in A and bAb\in A, and consider that there is a path γ:{0,,n}X\gamma:\{0,\ldots,n\}\rightarrow X such that γ(0)=a\gamma(0)=a and γ(n)=b\gamma(n)=b. Since AB=XA\cup B=X, there there exists ii such that γ(i)A\gamma(i)\in A and γ(i+1)B\gamma(i+1)\in B. Hence, there is λ:{0,1,2}X\lambda:\{0,1,2\}\rightarrow X in AABBA\sqcup_{A\cap B}B such that λ(0)=γ(i)\lambda(0)=\gamma(i) and λ(2)=γ(i+1)\lambda(2)=\gamma(i+1). By 3.4, we obtain that at least one of the three elements are in ABA\cap B ()(\rightarrow\leftarrow). Therefore, there is not a path between a,bXa,b\in X; XX is disconnected. ∎

4. Semi-coarse Strings

Building a groupoid in semi-coarse spaces demands some tricks and considerations. On one side, we would like to preserve the fundamental groups as a part of this groupoid, restricting to just one object and some morphisms. On the other hand, our motivation is to find and invariant which doesn’t lose all of the structure when we work in coarse spaces.

Therefore, we are going to explore bornologous maps f:1Xf:\mathbb{Z}_{1}\rightarrow X for a semi-coarse space XX. From this point, we call tails to the restrictions f[n,)f\mid_{[n,\infty)} and f(,m]f\mid_{(-\infty,m]} for some m,n,m,n,\in\mathbb{Z}. The tails are our main focus in the subsequent construction, and sooner we realize that it contains almost all of the information.

First, we give the main ingredients to create a groupoid: objects, morphism and a composition rule; for this construction we have symmetric maps, string and \star-operation, respectively.

Definition 4.1.

Let f:1Xf:\mathbb{Z}_{1}\rightarrow X:

  • ff is called symmetric if f(x)=f(x)f(x)=f(-x).

  • f¯\overline{f} is called the opposite direction of f:1Xf:\mathbb{Z}_{1}\rightarrow X with f¯(x)=f(x)\overline{f}(x)=f(-x)

Through the tails of the symmetric maps, we can develop a equivalence relation.

Definition 4.2.

Let f,gf,g be symmetric maps. ff and gg are eventually equal if there exists N,M0N,M\in\mathbb{N}_{0} such that f(N+k)=g(M+k)f(N+k)=g(M+k) for every k0k\in\mathbb{N}_{0}.

Proposition 4.3.

The relation of two symmetric maps being eventually equal is a equivalence relation. We write f\langle f\rangle_{\infty} to refer to the class of maps eventually equal to ff.

Taking the coarse space 1\mathbb{Z}_{1} as domain of our maps provides some structure in the image, that is, it compels the element f(z)f(z) to be related with f(z+1)f(z+1) and f(z1)f(z-1). However, being labeled by the integers proves to be really restrictive. The definition below unlabel these maps. This can be seen as being capable to move the functions in the integers axis and say they are exactly the same function. Later, that definition allows to obtain well defined constructions of merging and splitting these maps.

Definition 4.4.

Let f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X. ff and gg are called affine if there exists N0N\in\mathbb{N}_{0} such that f(x)=g(x+k)f(x)=g(x+k).

Affine is also an equivalence relation. We denote by fA{gf and g are affine}\langle f\rangle_{A}\coloneqq\{g\mid f\text{ and }g\text{ are affine}\}.

In 4.5, we are considering any map of the strings as affine classes.

Definition 4.5.

Let nn\in\mathbb{N}. We say that a F=(f1,,fn)F=(f_{1},\ldots,f_{n}) is an nn-string from f\langle f\rangle_{\infty} to g\langle g\rangle_{\infty} if we have the following:

  1. (1)

    f1f\overleftarrow{f_{1}}\in\langle f\rangle_{\infty} such that

    f1(x){f1(x)x0f1(x)x>0.\displaystyle\overleftarrow{f_{1}}(x)\coloneqq\left\{\begin{array}[]{ll}f_{1}(x)&x\leq 0\\ f_{1}(-x)&x>0.\end{array}\right.
  2. (2)

    fng\overrightarrow{f_{n}}\in\langle g\rangle_{\infty} such that

    fn(x){fn(x)x0fn(x)x>0.\displaystyle\overrightarrow{f_{n}}(x)\coloneqq\left\{\begin{array}[]{ll}f_{n}(x)&x\leq 0\\ f_{n}(-x)&x>0.\end{array}\right.
  3. (3)

    For every i{1,n1}i\in\{1,\ldots n-1\} there exists fi,RfiAf_{i,R}\in\langle f_{i}\rangle_{A} and fi+1,Lfi+1Af_{i+1,L}\in\langle f_{i+1}\rangle_{A} such that fi,R[0,)scfi+1,L¯[0,)f_{i,R}\mid_{[0,\infty)}\simeq_{sc}\overline{f_{i+1,L}}\mid_{[0,\infty)}.

Strings are the elements of the collection of all of the nn-strings for every nn\in\mathbb{N}.

We use the following notation in the subsequent sections, specially when we talk about the Van Kampen theorems:

  • X,A,n,f,g\mathcal{F}_{X,A,n,f,g} is the collection of all of the nn-string in XX from f\langle f\rangle_{\infty} to g\langle g\rangle_{\infty} such that their tails land in AXA\subset X.

  • If we omit the subset AA, we consider that A=XA=X.

  • If we omit the number nn we consider is all of the nn-strings with nn\in\mathbb{N}.

  • If we omit ff and gg, we consider that the strings can start and end in any element.

When the context is clear, we omit XX.

Definition 4.6.

Let F=(f1,,fn)F=(f_{1},\ldots,f_{n}) be a path from f\langle f\rangle_{\infty} to g\langle g\rangle_{\infty} and G=(g1,,gm)G=(g_{1},\ldots,g_{m}) be a path from g\langle g\rangle_{\infty} to h\langle h\rangle_{\infty}. We define (f1,,fn)(g1,gm)(f1,,fn,g1,,gm)(f_{1},\ldots,f_{n})\star(g_{1},\ldots g_{m})\coloneqq(f_{1},\ldots,f_{n},g_{1},\ldots,g_{m}).

F¯(fn¯,,f1¯)\overline{F}\coloneqq(\overline{f_{n}},\ldots,\overline{f_{1}}) is called the opposite direction of FF.

Although we define an operation between pair of strings, we are not ready to obtain a groupoid from this set. By this moment, it is not clear which element is the identity and how the product of an element with its inverse element produces such identity. To achieve that goal, we are interested in splitting and merging maps from 1\mathbb{Z}_{1} and deleting consecutive opposite maps.

The technical definition of this is divided in two relations are written in 4.7 and 4.13. The consecutive results and definitions are included to look for the correct way to merge maps and to see that this relation eventually build an equivalence relation.

Definition 4.7.

Let F=(f1,,fn)F=(f_{1},\ldots,f_{n})\in\mathcal{F}.

  • For i{2,,n2}i\in\{2,\ldots,n-2\}. If fifi+1¯Af_{i}\in\langle\overline{f_{i+1}}\rangle_{A}, then we say that the result of deleting two consecutive opposite maps in ii is F=(f1,,fi1,fi+2,,fn)F^{\prime}=(f_{1},\ldots,f_{i-1},f_{i+2},\ldots,f_{n}). We express this relation as Fdop(i)FF\xrightarrow{d_{op}(i)}F^{\prime}

  • For i{2,,n}i\in\{2,\ldots,n\}. Let gg we say that the result of adding the opposite maps gg and g¯\overline{g} in ii is F=(f1,,fi1,g,g¯,fi,,fn)F^{\prime}=(f_{1},\ldots,f_{i-1},g,\overline{g},f_{i},\ldots,f_{n}) if FF^{\prime}\in\mathcal{F}. We express this relation as Faop(i,g)FF\xrightarrow{a_{op}(i,g)}F^{\prime}.

Lemma 4.8.

Let FF\in\mathcal{F}. Then

  1. (1)

    If Fdop(i)FF\xrightarrow{d_{op}(i)}F^{\prime}, then FF^{\prime}\in\mathcal{F} and Faop(i,fi)FF^{\prime}\xrightarrow{a_{op}(i,f_{i})}F.

  2. (2)

    If Faop(i,g)FF\xrightarrow{a_{op}(i,g)}F^{\prime}, then Fdop(i)FF^{\prime}\xrightarrow{d_{op}(i)}F.

Proof.

Let FF\in\mathcal{F}.

(1) Suppose that Fdop(i)FF\xrightarrow{d_{op}(i)}F^{\prime}. By definition, fifi+1¯Af_{i}\in\langle\overline{f_{i+1}}\rangle_{A}. Without loss of generality, suppose that precisely fi=fi+1¯f_{i}=\overline{f_{i+1}}. In addition, there exists ri1,li,ri+1,li+2,D,Ur_{i-1},l_{i},r_{i+1},l_{i+2},D,U\in\mathbb{Z} and H1,H2:[0,)×1XH_{1},H_{2}:[0,\infty)\times\mathbb{Z}_{1}\rightarrow X such that

H1(z,k)=\displaystyle H_{1}(z,k)\ =\ fi1(ri1+z) if kD\displaystyle f_{i-1}(r_{i-1}+z)\text{ if }k\leq D
H1(z,k)=\displaystyle H_{1}(z,k)\ =\ fi(liz) if k0\displaystyle f_{i}(l_{i}-z)\text{ if }k\geq 0
H2(z,k)=\displaystyle H_{2}(z,k)\ =\ fi+1(ri+z) if k0\displaystyle f_{i+1}(r_{i}+z)\text{ if }k\leq 0
H2(z,k)=\displaystyle H_{2}(z,k)\ =\ fi+2(li+1z) if kU.\displaystyle f_{i+2}(l_{i+1}-z)\text{ if }k\geq U.

Since fi=fi+1¯f_{i}=\overline{f_{i+1}}, fi+1(zli)=fi(liz)f_{i+1}(z-l_{i})=f_{i}(l_{i}-z). Take amax{ri,li}a\coloneqq\max\{r_{i},-l_{i}\}. Displace H1H_{1} and H2H_{2} such that

H1(z,k)=\displaystyle H^{\prime}_{1}(z,k)\ =\ fi1(ri1+|a+li|+z) if kD\displaystyle f_{i-1}(r_{i-1}+|a+l_{i}|+z)\text{ if }k\leq D
H1(z,k)=\displaystyle H^{\prime}_{1}(z,k)\ =\ fi(li|a+li|z) if k0\displaystyle f_{i}(l_{i}-|a+l_{i}|-z)\text{ if }k\geq 0
H2(z,k)=\displaystyle H^{\prime}_{2}(z,k)\ =\ fi+1(ri+|ria|+z) if k0\displaystyle f_{i+1}(r_{i}+|r_{i}-a|+z)\text{ if }k\leq 0
H2(z,k)=\displaystyle H^{\prime}_{2}(z,k)\ =\ fi+1(li+1|ria|z) if kU.\displaystyle f_{i+1}(l_{i+1}-|r_{i}-a|-z)\text{ if }k\geq U.

Therefore, we built a homotopy from some fi1,Rfi1Af^{\prime}_{i-1,R}\in\langle f_{i-1}\rangle_{A} to some fi+2,Lfi+1Af^{\prime}_{i+2,L}\in\langle f_{i+1}\rangle_{A}. Thus FF^{\prime}\in\mathcal{F}.

On the other hand, since Fdop(i)FF\xrightarrow{d_{op}(i)}F^{\prime}, then FF\in\mathcal{F} and fifi+1¯Af_{i}\in\langle\overline{f_{i+1}}\rangle_{A}, thus Faop(i,g)FF^{\prime}\xrightarrow{a_{op}(i,g)}F.

(2) Let Fas(i,g)FF\xrightarrow{a_{s}(i,g)}F^{\prime}. By definition Fds(i)FF^{\prime}\xrightarrow{d_{s}(i)}F because the element after gg in FF^{\prime} is precisely g¯\overline{g}. ∎

Definition 4.9.

Let f:Xf:\mathbb{Z}\rightarrow X be a map.

  • ff is called periodic if there exists TT\in\mathbb{N} such that f(x)=f(x+T)f(x)=f(x+T) for every xXx\in X.

  • ff is called eventually right periodic if there exist TT\in\mathbb{N} and zz\in\mathbb{Z} such that f(x)=f(x+T)f(x)=f(x+T) for every xzx\geq z.

  • ff is called eventually left periodic if there exists TT\in\mathbb{N} and zz\in\mathbb{Z} such that f(x)=f(xT)f(x)=f(x-T) for evry xzx\leq z.

Lemma 4.10.

Let f:Xf:\mathbb{Z}\rightarrow X be a map.

  1. (1)

    If ff is periodic, then ff is both eventually right and left periodic.

  2. (2)

    If ff is not periodic and is eventually right periodic, then

    min{zf(x)=f(x+T) for every xz}.\displaystyle\min\{z\in\mathbb{Z}\mid f(x)=f(x+T)\text{ for every }x\geq z\}\in\mathbb{Z}.
  3. (3)

    If ff is not periodic and is eventually left periodic, then

    max{zf(x)=f(xT) for every xz}.\displaystyle\max\{z\in\mathbb{Z}\mid f(x)=f(x-T)\text{ for every }x\leq z\}\in\mathbb{Z}.
Proof.

Let f:Xf:\mathbb{Z}\rightarrow X be a map.

(1) Let ff be periodic, then there exists TT\in\mathbb{N} such that f(x)=f(x+T)f(x)=f(x+T) for every xXx\in X. Thus f(x)=f(xT)f(x)=f(x-T) for every xXx\in X. In particular, f(x)=f(x+T)f(x)=f(x+T) for every x0x\geq 0 and f(y)=f(yT)f(y)=f(y-T) for every y0y\leq 0.

(2) Let ff be eventually right periodic and not periodic. Then

A\displaystyle A\coloneqq {zf(x)=f(x+T) for every xz} and\displaystyle\{z\in\mathbb{Z}\mid f(x)=f(x+T)\text{ for every }x\geq z\}\in\mathbb{Z}\neq\varnothing\text{ and }
z\displaystyle z\in Az+1A.\displaystyle A\Rightarrow z+1\in A.

If minA\min\ A\notin\mathbb{Z}, then f(x)=f(x+T)f(x)=f(x+T) for every xx\in\mathbb{Z} ()(\rightarrow\leftarrow). Thus minA\min\ A\in\mathbb{Z}.

(3) Let ff be eventually left periodic and not periodic. Then

A\displaystyle A\coloneqq {zf(x)=f(xT) for every xz} and\displaystyle\{z\in\mathbb{Z}\mid f(x)=f(x-T)\text{ for every }x\leq z\}\in\mathbb{Z}\neq\varnothing\text{ and }
z\displaystyle z\in Az1A.\displaystyle A\Rightarrow z-1\in A.

If maxA\max\ A\notin\mathbb{Z}, then f(x)=f(xT)f(x)=f(x-T) for every xx\in\mathbb{Z} ()(\rightarrow\leftarrow). Thus maxA\max\ A\in\mathbb{Z}. ∎

Lemma 4.11.

Let f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X be not periodic bornologous maps such that there exists a,b,a,ba,b,a^{\prime},b^{\prime}\in\mathbb{Z} such that f(a+z)=g(bz)f(a+z)=g(b-z) and f(a+z)=g(bz)f(a^{\prime}+z)=g(b^{\prime}-z) for every z0z\geq 0. Then aa=bba-a^{\prime}=b^{\prime}-b or gg is eventually left periodic.

Proof.

Without loss of generality, suppose that aaa\leq a^{\prime}, and consider d|bb|d\coloneqq|b^{\prime}-b|. Then, aa+d0a^{\prime}-a+d\geq 0.

By hypothesis , we obtain that

f(a+(aa+d+z))=\displaystyle f(a+(a^{\prime}-a+d+z))= g(b(aa+d+z)) and\displaystyle g(b-(a^{\prime}-a+d+z))\text{ and }
f(a+(aa+d+z))=\displaystyle f(a+(a^{\prime}-a+d+z))= f(a+(d+z))=g(b(d+z)).\displaystyle f(a^{\prime}+(d+z))=g(b^{\prime}-(d+z)).

for every z0z\geq 0.

On one hand, suppose that bb0b^{\prime}-b\geq 0. Then

g(b(d+z))=g(b(bb+z))=g(bz).\displaystyle g(b^{\prime}-(d+z))=g(b^{\prime}-(b^{\prime}-b+z))=g(b-z).

Thus, g(bTz)=g(bz)g(b-T-z)=g(b-z) for every z0z\geq 0 with T=(aa)+dT=(a^{\prime}-a)+d.

On the other hand, suppose that bb0b-b^{\prime}\geq 0. Then

g(b(aa+d+z))=\displaystyle g(b-(a^{\prime}-a+d+z))= g(b(aa+z)b+b)=g(b(aa+z))\displaystyle g(b-(a^{\prime}-a+z)-b+b^{\prime})=g(b^{\prime}-(a^{\prime}-a+z))
g(b(d+z))=\displaystyle g(b^{\prime}-(d+z))= g(b(bb+z))\displaystyle g(b^{\prime}-(b-b^{\prime}+z))

If bb=aab-b^{\prime}=a^{\prime}-a, we cannot ensure any about the periodicity. If they are different, we can define m=max{bb,aa}m=\max\{b-b^{\prime},a^{\prime}-a\} and see that g(bmTz)=g(bmz)g(b^{\prime}-m-T-z)=g(b^{\prime}-m-z) for every z0z\geq 0 and with T=|aa+bb|T=|a^{\prime}-a+b^{\prime}-b|. In both cases gg is eventually left cyclic or bb=aab-b^{\prime}=a^{\prime}-a. ∎

Proposition 4.12.

Let f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X be not periodic bornologous maps such that fg¯Af\notin\langle\overline{g}\rangle_{A} and there exist a,ba,b\in\mathbb{Z} such that f(a+z)=g(bz)f(a+z)=g(b-z) for every z0z\geq 0. There exists unique a,ba^{\prime},b^{\prime}\in\mathbb{Z} such that f(a+z)=g(bz)f(a^{\prime}+z)=g(b^{\prime}-z) for every z0z\geq 0 and f(a1)g(b+1)f(a^{\prime}-1)\neq g(b^{\prime}+1).

Proof.

We divide this proof in two case, whether ff is eventually right periodic. We can observe that ff is eventually right periodic if and only if gg is eventually left periodic, so it’s enough to consider these two cases.

Suppose that ff is not eventually right periodic. If a,ba^{\prime},b^{\prime}\in\mathbb{Z} satisfy that f(a+z)=g(bz)f(a^{\prime}+z)=g(b^{\prime}-z) for every z0z\geq 0, then, by 4.11, we observe that aa=bba^{\prime}-a=b-b^{\prime}; therefore we can write it as f(a+z)=g(bz)f(a+z)=g(b-z) for every zaaz\geq a^{\prime}-a. Since fg¯Af\notin\langle\overline{g}\rangle_{A} we have that there exists m<0m<0 such that f(a+m)=g(b+m)f(a+m)=g(b+m). Then,

M{zf(a+z)=g(bz) for every zz}\displaystyle M\coloneqq\{z^{\prime}\in\mathbb{Z}\mid f(a+z)=g(b-z)\text{ for every }z\geq z^{\prime}\}\in\mathbb{Z}

Thus, f(a+M+z)=g(bMz)f(a+M+z)=g(b-M-z) for every z0z\geq 0 and f(a+M1)g(bM+1)f(a+M-1)\neq g(b-M+1).

Suppose that ff is eventually right periodic with period TT, then gg is eventually left periodic with period TT. Applying 4.10, there exists A,BA,B\in\mathbb{Z} such that

A=\displaystyle A= min{zf(x)=f(x+T) for every xz}\displaystyle\min\{z\in\mathbb{Z}\mid f(x)=f(x+T)\text{ for every }x\geq z\}
B=\displaystyle B= max{zg(x)=g(xT) for every xz}\displaystyle\max\{z\in\mathbb{Z}\mid g(x)=g(x-T)\text{ for every }x\leq z\}

First we work with the case aAa\leq A. Observe that f(A1)g(A+T1)f(A-1)\neq g(A+T-1). If f(A1)=g(bA+a+1)f(A-1)=g(b-A+a+1), then g(bA+a+1)g(bA+a+1T)g(b-A+a+1)\neq g(b-A+a+1-T), and g(bA+az)=g(bA+aTz)g(b-A+a-z)=g(b-A+a-T-z) for every z0z\geq 0; thus bA+a+1=B+1b-A+a+1=B+1. Hence bB=Aa0b-B=A-a\geq 0. We obtain that

f(A)=f(a+(Aa))=g(b(Aa))=g(b(bB))=g(B).\displaystyle f(A)=f(a+(A-a))=g(b-(A-a))=g(b-(b-B))=g(B).

The procedure is completely analogous for not eventually right periodic, obtaining analogous values aa^{\prime} and bb^{\prime}.

On the other hand, if f(A1)g(bA+a+1)f(A-1)\neq g(b-A+a+1), then aAa\geq A. Hence a=Aa=A and BbB\geq b. We consider

bmax{bf(A+z)=g(bz) for every z0},\displaystyle b^{\prime}\coloneqq\max\{b^{\prime}\in\mathbb{Z}\mid f(A+z)=g(b^{\prime}-z)\text{ for every }z\geq 0\},

observing that bbb^{\prime}\geq b. Hence, for this case, aa=Aa^{\prime}\coloneqq a=A and bb^{\prime} is the element we have found.

We now consider that a>Aa>A, which implies that BbB\geq b, because at least g(b)=g(bT)g(b)=g(b-T). Define d{aA,Bb}d\coloneqq\{a-A,B-b\}, then f(ad+z)=g(b+dz)f(a-d+z)=g(b+d-z) by periodicity.

Quickly we observe that:

  • If aA=Bba-A=B-b, then f(A+z)=g(Bz)f(A+z)=g(B-z) and we proceed in similar way as not eventually right periodic.

  • If aA<Bba-A<B-b, we compute

    amin{af(ad+z)=g(b+dz) for every z0}\displaystyle a^{\prime}\coloneqq\min\{a^{\prime}\in\mathbb{Z}\mid f(a^{\prime}-d+z)=g(b+d-z)\text{ for every }z\geq 0\}

    and the values we want to get was ada^{\prime}-d and b+db+d.

  • If aA>Bba-A>B-b, we compute.

    bmin{bf(ad+z)=g(b+dz) for every z0}\displaystyle b^{\prime}\coloneqq\min\{b^{\prime}\in\mathbb{Z}\mid f(a-d+z)=g(b^{\prime}+d-z)\text{ for every }z\geq 0\}

    and the values we want to get was ada-d and b+db^{\prime}+d.∎

For a technical reason, we are just going to merge maps which are neither eventually right periodic nor eventually left periodic, with the exception of period 1, 2 or 3.

Definition 4.13.

Let F=(f1,,fn)F=(f_{1},\ldots,f_{n})\in\mathcal{F}. For i{1,,n1}i\in\{1,\ldots,n-1\}, fifi+1¯Af_{i}\notin\langle\overline{f_{i+1}}\rangle_{A} and there exists a,ba,b\in\mathbb{Z} such that fi(a+z)=fi+1(bz)f_{i}(a+z)=f_{i+1}(b-z) for every z0z\geq 0, we say that the result of merging fif_{i} and fi+1f_{i+1} is F=(f1,,fi1,g,fi+2,fn)F^{\prime}=(f_{1},\ldots,f_{i-1},g,f_{i+2},\ldots f_{n}) such that

  • If fif_{i} is periodic with T{1,2,3}T\in\{1,2,3\} and fi+1f_{i+1} is not periodic, then by 4.10 there exists bb^{\prime} such fi+1(x)=fi+1(xT)f_{i+1}(x)=f_{i+1}(x-T) for every xbx\leq b^{\prime} and fi+1(b+1)fi+1(b+1T)f_{i+1}(b+1)\neq f_{i+1}(b+1-T), and

    g(z){fi(a+z) if z0fi+1(b+z) if z>0.\displaystyle g(z)\coloneqq\left\{\begin{array}[]{ll}f_{i}(a+z)&\text{ if }z\leq 0\\ f_{i+1}(b^{\prime}+z)&\text{ if }z>0.\end{array}\right.
  • If fi+1f_{i+1} is periodic with T{1,2,3}T\in\{1,2,3\} and fif_{i} is not periodic, then by 4.10 there exists aa^{\prime} such f(x)=f(x+T)f(x)=f(x+T) for every xbx\geq b^{\prime} and f(a1)=f(a1+T)f(a-1)=f(a-1+T), and

    g(z){fi(a+z) if z0fi+1(b+z) if z>0.\displaystyle g(z)\coloneqq\left\{\begin{array}[]{ll}f_{i}(a^{\prime}+z)&\text{ if }z\leq 0\\ f_{i+1}(b+z)&\text{ if }z>0.\end{array}\right.
  • If fi,fi+1f_{i},f_{i+1} are not periodic, and fif_{i} is eventually right periodic with T{1,2,3}T\in\{1,2,3\} or is not eventually right periodic. Applying 4.12, we find a,ba^{\prime},b^{\prime}\in\mathbb{Z} such that fi(a+z)=fi+1(bz)f_{i}(a^{\prime}+z)=f_{i+1}(b^{\prime}-z) for every zz\in\mathbb{Z} and fi(a1)fi+1(b+1)f_{i}(a^{\prime}-1)\neq f_{i+1}(b^{\prime}+1). We define

    g(z){fi(a+z) if z0fi+1(b+z) if z>0.\displaystyle g(z)\coloneqq\left\{\begin{array}[]{ll}f_{i}(a^{\prime}+z)&\text{ if }z\leq 0\\ f_{i+1}(b^{\prime}+z)&\text{ if }z>0.\end{array}\right.

On the other hand, if fifi+1¯Af_{i}\in\langle\overline{f_{i+1}}\rangle_{A} and fif_{i} is eventually left periodic with T{1,2}T\in\{1,2\} we define g(2)=fi(a)g(2\mathbb{Z})=f_{i}(a) and g(21)=fi(a1)g(2\mathbb{Z}-1)=f_{i}(a-1) where aa satisfies that fi(x)=fi(xT)f_{i}(x)=f_{i}(x-T) for every xax\leq a.

We express this relation as Fm(i)FF\xrightarrow{m(i)}F^{\prime} (see an exemplification in Figure 3).

fif_{i}fi+1f_{i+1}m(i)m(i)
Figure 3. Merging fif_{i} and fi+1f_{i+1} which are maps from 2\mathbb{Z}_{2} to 2\mathbb{R}^{2}. The darker the color in the point, the higher its value in the integers.

For i{1,,n}i\in\{1,\ldots,n\}, we sat that the result of split fif_{i} in gg and hh at ziz_{i} is F=(f1,,fi1,g,h,fi+1,,fn)F^{\prime}=(f_{1},\ldots,f_{i-1},g,h,f_{i+1},\ldots,f_{n}) is a string with:

  • If fif_{i} is periodic with T{1,2}T\in\{1,2\} and there exist m,nm,n\in\mathbb{Z} such that

    g(m+k)=\displaystyle g(m+k)=\ h(nk)k,\displaystyle h(n-k)\ \forall k\in\mathbb{Z},
    g(m+k)=\displaystyle g(m+k)=\ fi(k)k0.\displaystyle f_{i}(k)\ \forall k\leq 0.
  • If there exists zz\in\mathbb{Z} such that fi(z1)fi(z+1)f_{i}(z-1)\neq f_{i}(z+1), take ziz_{i}\in\mathbb{Z} with that condition and there exist m,nm,n\in\mathbb{Z} such that

    g(m+k)=\displaystyle g(m+k)=\ h(nk)k0,\displaystyle h(n-k)\ \forall k\geq 0,
    g(m+k)=\displaystyle g(m+k)=\ fi(zi+k)k0,\displaystyle f_{i}(z_{i}+k)\ \forall k\leq 0,
    h(n+k)=\displaystyle h(n+k)=\ fi(zi+k)k0.\displaystyle f_{i}(z_{i}+k)\ \forall k\geq 0.

where gg is eventually right periodic with T{1,2,3}T\in\{1,2,3\} or is not eventually right periodic. We express this relation as Fs(i,g,h)FF\xrightarrow{s(i,g,h)}F^{\prime} (see an exemplification in Figure 3).

fif_{i}gghh
Figure 4. Splitting fif_{i}, a map from 2\mathbb{Z}_{2} to 2\mathbb{R}^{2}, in gg and hh.
Lemma 4.14.

Let F=(f1,,fn)F=(f_{1},\ldots,f_{n})\in\mathcal{F}. Then

  1. (1)

    m(i)m(i) is well-defined.

  2. (2)

    If Fm(i)FF\xrightarrow{m(i)}F^{\prime}, then Fs(i,fi,fi+1)FF^{\prime}\xrightarrow{s(i,f_{i},f_{i+1})}F.

  3. (3)

    If Fs(i,g,h)FF\xrightarrow{s(i,g,h)}F^{\prime}, then Fm(i)FF^{\prime}\xrightarrow{m(i)}F.

Proof.

Let FF\in\mathcal{F}.

(1) Suppose that for i{1,,n1}i\in\{1,\ldots,n-1\} we have that Fm(i)FF\xrightarrow{m(i)}F^{\prime}. Then fifi+1¯Af_{i}\notin\langle\overline{f_{i+1}}\rangle_{A} and there exists a,ba,b\in\mathbb{Z} such that fi(a+z)=fi+1(bz)f_{i}(a+z)=f_{i+1}(b-z) for every z0z\geq 0. Hence FF^{\prime} is a string because g(z)=fi(a+z)g(z)=f_{i}(a+z) for every z0z\leq 0 and g(z)=fi+1(b+z)g(z)=f_{i+1}(b+z) for every z>0z>0.

  • Suppose that fif_{i} is periodic with T{1,2,3}T\in\{1,2,3\} and fi+1f_{i+1} is not periodic. Consider a′′a^{\prime\prime}\in\mathbb{Z} such that fi(a′′+z)=fi+1(bz)f_{i}(a^{\prime\prime}+z)=f_{i+1}(b^{\prime}-z) for every zz\in\mathbb{Z}, then T=|aa′′|T=|a-a^{\prime\prime}| we define

    g′′(z){fi(a′′+z) if z0fi+1(b+z) if z>0.\displaystyle g^{\prime\prime}(z)\coloneqq\left\{\begin{array}[]{ll}f_{i}(a^{\prime\prime}+z)&\text{ if }z\leq 0\\ f_{i+1}(b^{\prime}+z)&\text{ if }z>0.\end{array}\right.

    Since fi(a+z)=fi(a′′(a′′a)+z)=fi(a′′+z)f_{i}(a+z)=f_{i}(a^{\prime\prime}-(a^{\prime\prime}-a)+z)=f_{i}(a^{\prime\prime}+z). Thus g=gg=g^{\prime}.

  • Suppose that fi+1f_{i+1} is periodic with T{1,2,3}T\in\{1,2,3\} and fif_{i} is not periodic. In analogous way of the case above, we can select any a,ba,b\in\mathbb{Z} such that fi(a+z)=fi+1(bz)f_{i}(a+z)=f_{i+1}(b-z) for every zz\in\mathbb{Z}.

  • Suppose that fi,fi+1f_{i},f_{i+1} are not periodic, and fif_{i} is eventually right periodic with T{1,2,3}T\in\{1,2,3\} or is not eventually right periodic. Since a,ba^{\prime},b^{\prime} are unique, then m(i)m(i) is well defined.

(2) Suppose that Fm(i)FF\xrightarrow{m(i)}F^{\prime}. We explore the whole four cases, just in the first three cases we consider that fifi+1Af_{i}\notin\langle f_{i+1}\rangle_{A}:

  • Suppose that fif_{i} is periodic with T{1,2,3}T\in\{1,2,3\} and fi+1f_{i+1} is not periodic. By 4.10, there exists bb^{\prime} such that fi+1(x)=fi+1(xT)f_{i+1}(x)=f_{i+1}(x-T) for every xbx\leq b^{\prime} and f(b+1)f(b+1T)f(b+1)\neq f(b+1-T). Hence, building gg as in 4.13, g(1)g(1)g(-1)\neq g(1) and by definition

    fi(a+k)=\displaystyle f_{i}(a+k)= fi+1(bk) for every k0,\displaystyle f_{i+1}(b^{\prime}-k)\text{ for every }k\geq 0,
    fi(a+k)=\displaystyle f_{i}(a+k)= g(k) for every k0,\displaystyle g(k)\text{ for every }k\leq 0,
    fi+1(b+k)=\displaystyle f_{i+1}(b^{\prime}+k)= g(k) for every k0.\displaystyle g(k)\text{ for every }k\geq 0.

    Thus Fs(0,fi,fi+1)FF^{\prime}\xrightarrow{s(0,f_{i},f_{i+1})}F.

  • Suppose that fi+1f_{i+1} is periodic with T{1,2,3}T\in\{1,2,3\} and fif_{i} is not periodic. The case analogous but we apply 4.10 in fif_{i}. Thus Fs(0,fi,fi+1)FF^{\prime}\xrightarrow{s(0,f_{i},f_{i+1})}F.

  • Suppose that fi,fi+1f_{i},f_{i+1} are not periodic, and fif_{i} is eventually right periodic with T{1,2,3}T\in\{1,2,3\} or is not eventually right periodic. Then there exist a,ba^{\prime},b^{\prime}\in\mathbb{Z} such that fi(a+z)=fi+1(bz)f_{i}(a^{\prime}+z)=f_{i+1}(b^{\prime}-z) for every z0z\geq 0 and fi(a1)fi+1(b+1)f_{i}(a^{\prime}-1)\neq f_{i+1}(b^{\prime}+1). Hence, building gg as in 4.13, g(1)g(1)g(-1)\neq g(1) and by definition

    fi(a+k)=\displaystyle f_{i}(a^{\prime}+k)= fi+1(bk) for every k0,\displaystyle f_{i+1}(b^{\prime}-k)\text{ for every }k\geq 0,
    fi(a+k)=\displaystyle f_{i}(a^{\prime}+k)= g(k) for every k0,\displaystyle g(k)\text{ for every }k\leq 0,
    fi+1(b+k)=\displaystyle f_{i+1}(b^{\prime}+k)= g(k) for every k0.\displaystyle g(k)\text{ for every }k\geq 0.

    Thus Fs(i,fi,fi+1)FF^{\prime}\xrightarrow{s(i,f_{i},f_{i+1})}F.

  • Suppose that fifi+1¯Af_{i}\in\langle\overline{f_{i+1}}\rangle_{A} and fif_{i} is eventually left periodic with T{1,2}T\in\{1,2\}. Then there exist a,ba,b\in\mathbb{Z} such that fi(a+z)=fi+1(bz)f_{i}(a+z)=f_{i+1}(b-z) for every zz\in\mathbb{Z} and there exists aa^{\prime}\in\mathbb{Z} such that fi(z)=fi(zT)f_{i}(z)=f_{i}(z-T) for every zaz\leq a^{\prime} and fi(a+1)fi(a+1T)f_{i}(a^{\prime}+1)\neq f_{i}(a^{\prime}+1-T), and g(2)=fi(a)g(2\mathbb{Z})=f_{i}(a^{\prime}) and g(2+1)=fi(a1)g(2\mathbb{Z}+1)=f_{i}(a^{\prime}-1). Hence,

    fi(a+k)=\displaystyle f_{i}(a^{\prime}+k)= fi+1(b(aa)k) for every k,\displaystyle f_{i+1}(b^{\prime}-(a^{\prime}-a)-k)\text{ for every }k\in\mathbb{Z},
    fi(a+k)=\displaystyle f_{i}(a^{\prime}+k)= g(k) for every k0.\displaystyle g(k)\text{ for every }k\leq 0.

    Thus Fs(i,fi,fi+1)FF^{\prime}\xrightarrow{s(i,f_{i},f_{i+1})}F.

(3) Suppose that Fs(i,g,h)FF\xrightarrow{s(i,g,h)}F^{\prime}. We consider both of the possible splittings.

  • Suppose that fif_{i} is periodic with T{1,2}T\in\{1,2\} and there exist m,nm,n\in\mathbb{Z} such that g(m+k)=h(nk)g(m+k)=h(n-k) for every kk\in\mathbb{Z} and g(m+k)=fi(k)g(m+k)=f_{i}(k) for every k0k\leq 0. Then we have that gh¯Ag\in\langle\overline{h}\rangle_{A} and gg is eventually left periodic. Therefore we can merge them an create gfiAg^{\prime}\in\langle f_{i}\rangle_{A}. Thus, Fm(i)FF^{\prime}\xrightarrow{m(i)}F.

  • Suppose that there exists z0z_{0}\in\mathbb{Z} such that fi(z01)fi(z0+1)f_{i}(z_{0}-1)\neq f_{i}(z_{0}+1), and there exist m,nm,n\in\mathbb{Z} such that g(m+k)=h(nk)g(m+k)=h(n-k) for every k0k\geq 0, g(m+k)=fi(z0+k)g(m+k)=f_{i}(z_{0}+k) for every k0k\leq 0, and h(n+k)=fi(z0+k)h(n+k)=f_{i}(z_{0}+k) for every k0k\geq 0. Observe that h(n+1)g(m1)h(n+1)\neq g(m-1). Then, we might apply any of the three possibilities to merge. Thus Fm(i)FF^{\prime}\xrightarrow{m(i)}F.∎

Definition 4.15.

Let F,GF,G\in\mathcal{F}. We say that we can convert FF in GG if there exists a finite sequence of arrows

FF0α0F1α1αn1FnG\displaystyle F\coloneqq F_{0}\xrightarrow{\alpha_{0}}F_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}F_{n}\eqqcolon G

with αi{dop(i)i}{aop(i,g)i,g:1X}{m(i)i}{s(i,g,h)i,g,h:1X}\alpha_{i}\in\{d_{op}(i)\mid i\in\mathbb{Z}\}\cup\{a_{op}(i,g)\mid i\in\mathbb{Z},\ g:\mathbb{Z}_{1}\rightarrow X\}\cup\{m(i)\mid i\in\mathbb{Z}\}\cup\{s(i,g,h)\mid i\in\mathbb{Z},\ g,h:\mathbb{Z}_{1}\rightarrow X\}. We express this relation as FSGF\simeq_{S}G.

Proposition 4.16.

S\simeq_{S} is an equivalence relation.

Proof.

Let F,G,HF,G,H\in\mathcal{F} with F=(f1,,fn)F=(f_{1},\ldots,f_{n}).

Reflexive: In a first case, consider that there exist z1z_{1}\in\mathbb{Z} such that f1(z11)f1(z1+1)f_{1}(z_{1}-1)\neq f_{1}(z_{1}+1). Define

g{f1(z1+z) if z0f1(z1) if z>0,h{f1(z1+z) if z0f1(z1) if z<0.\displaystyle g\coloneqq\left\{\begin{array}[]{ll}f_{1}(z_{1}+z)&\text{ if }z\leq 0\\ f_{1}(z_{1})&\text{ if }z>0\end{array}\right.,\ h\coloneqq\left\{\begin{array}[]{ll}f_{1}(z_{1}+z)&\text{ if }z\geq 0\\ f_{1}(z_{1})&\text{ if }z<0.\end{array}\right.

On the other hand, for every zz\in\mathbb{Z}, f1(z1)=f1(z)f_{1}(z-1)=f_{1}(z) and f1f_{1} is periodic with T{1,2}T\in\{1,2\}. Define

g{f1(z) if z0f1(0) if z>0,h{f1(z) if z0f1(0) if z<0.\displaystyle g\coloneqq\left\{\begin{array}[]{ll}f_{1}(z)&\text{ if }z\leq 0\\ f_{1}(0)&\text{ if }z>0\end{array}\right.,\ h\coloneqq\left\{\begin{array}[]{ll}f_{1}(z)&\text{ if }z\geq 0\\ f_{1}(0)&\text{ if }z<0.\end{array}\right.

In both case we have that Fs(1,g,h)(g,h,f2,)m(1)FF\xrightarrow{s(1,g,h)}(g,h,f_{2},\ldots)\xrightarrow{m(1)}F.

Symmetry: Suppose that FSGF\simeq_{S}G. Then, there exist the sequence

FF0α0F1α1αn1FnG\displaystyle F\coloneqq F_{0}\xrightarrow{\alpha_{0}}F_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}F_{n}\eqqcolon G

with αi{dop(i)i}{aop(i,g)i,g:1X}{m(i)i}{s(i,g,h)i,g,h:1X}\alpha_{i}\in\{d_{op}(i)\mid i\in\mathbb{Z}\}\cup\{a_{op}(i,g)\mid i\in\mathbb{Z},\ g:\mathbb{Z}_{1}\rightarrow X\}\cup\{m(i)\mid i\in\mathbb{Z}\}\cup\{s(i,g,h)\mid i\in\mathbb{Z},\ g,h:\mathbb{Z}_{1}\rightarrow X\}. By 4.8 and 4.14, we know that every map αi\alpha_{i} have his opposite arrow βi{dop(i)i}{aop(i,g)i,g:1X}{m(i)i}{s(i,g,h)i,g,h:1X}\beta_{i}\in\{d_{op}(i)\mid i\in\mathbb{Z}\}\cup\{a_{op}(i,g)\mid i\in\mathbb{Z},\ g:\mathbb{Z}_{1}\rightarrow X\}\cup\{m(i)\mid i\in\mathbb{Z}\}\cup\{s(i,g,h)\mid i\in\mathbb{Z},\ g,h:\mathbb{Z}_{1}\rightarrow X\} such that

Fnβn1Fn1αn2α0F0\displaystyle F_{n}\xrightarrow{\beta_{n-1}}F_{n-1}\xrightarrow{\alpha_{n-2}}\cdots\xrightarrow{\alpha_{0}}F_{0}

Hence GSFG\simeq_{S}F.

Transitive: Suppose that FSGF\simeq_{S}G and GSHG\simeq_{S}H. Then, there exist the sequence

FF0α0F1α1αn1FnG\displaystyle F\coloneqq F_{0}\xrightarrow{\alpha_{0}}F_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}F_{n}\eqqcolon G
GG0β0G1β1βn1GnH\displaystyle G\coloneqq G_{0}\xrightarrow{\beta_{0}}G_{1}\xrightarrow{\beta_{1}}\cdots\xrightarrow{\beta_{n-1}}G_{n}\eqqcolon H

with αi{dop(i)i}{aop(i,g)i,g:1X}{m(i)i}{s(i,g,h)i,g,h:1X}\alpha_{i}\in\{d_{op}(i)\mid i\in\mathbb{Z}\}\cup\{a_{op}(i,g)\mid i\in\mathbb{Z},\ g:\mathbb{Z}_{1}\rightarrow X\}\cup\{m(i)\mid i\in\mathbb{Z}\}\cup\{s(i,g,h)\mid i\in\mathbb{Z},\ g,h:\mathbb{Z}_{1}\rightarrow X\}. Therefore

FF0α0F1α1αn1Fn=G=G0β0G1β1βn1GnH\displaystyle F\coloneqq F_{0}\xrightarrow{\alpha_{0}}F_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}F_{n}=G=G_{0}\xrightarrow{\beta_{0}}G_{1}\xrightarrow{\beta_{1}}\cdots\xrightarrow{\beta_{n-1}}G_{n}\eqqcolon H

and thus FSHF\simeq_{S}H. ∎

Let F=(f1,,fn)F=(f_{1},\ldots,f_{n})\in\mathcal{F}, we denote the class of FF as [F]S[F]_{S} or [f1,,fn]S[f_{1},\ldots,f_{n}]_{S}. Finally, we develop the operation between classes.

Proposition 4.17.

Let F,GF,G\in\mathcal{F}. Then [F]S[G]S=[FG]S[F]_{S}\star[G]_{S}=[F\star G]_{S} is well-defined.

Proof.

Let F,GF,G\in\mathcal{F}, F[F]SF^{\prime}\in[F]_{S} and G[G]SG^{\prime}\in[G]_{S}, then there exist the sequences

FF0α0F1α1αn1FnF,\displaystyle F\coloneqq F_{0}\xrightarrow{\alpha_{0}}F_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}F_{n}\eqqcolon F^{\prime},
GG0β0G1β1βn1GmG.\displaystyle G\coloneqq G_{0}\xrightarrow{\beta_{0}}G_{1}\xrightarrow{\beta_{1}}\cdots\xrightarrow{\beta_{n-1}}G_{m}\eqqcolon G^{\prime}.

Then we have that

FG=\displaystyle F\star G= F0Gα01GF1Gα11Gαn11GFnG=FG,\displaystyle F_{0}\star G\xrightarrow{\alpha_{0}\star 1_{G}}F_{1}\star G\xrightarrow{\alpha_{1}\star 1_{G}}\cdots\xrightarrow{\alpha_{n-1}\star 1_{G}}F_{n}\star G=F^{\prime}\star G,
FG=\displaystyle F^{\prime}\star G= FG01Fβ0FG11Fβ11Fβm1FGm=FG.\displaystyle F^{\prime}\star G_{0}\xrightarrow{1_{F^{\prime}}\star\beta_{0}}F^{\prime}\star G_{1}\xrightarrow{1_{F^{\prime}}\star\beta_{1}}\cdots\xrightarrow{1_{F^{\prime}}\star\beta_{m-1}}F^{\prime}\star G_{m}=F^{\prime}\star G^{\prime}.\qed

.

Considering this class, we observe in the following lemma that with this operation every element has inverses and left and right identities. Hence we are able to define the groupoid.

Lemma 4.18.

Let F=(f1,,fn)F=(f_{1},\ldots,f_{n})\in\mathcal{F}, then [F]S[F¯]S=[f1,f1¯]S[F]_{S}\star[\overline{F}]_{S}=[f_{1},\overline{f_{1}}]_{S} and [F¯]S[F]S=[fn¯,fn]S[\overline{F}]_{S}\star[F]_{S}=[\overline{f_{n}},f_{n}]_{S}. Moreover, [f1,f1¯]S[F]S=[F]S=[F]S[fn¯,fn]S[f_{1},\overline{f_{1}}]_{S}\star[F]_{S}=[F]_{S}=[F]_{S}\star[\overline{f_{n}},f_{n}]_{S}.

Proof.

Let F=(f1,,fn)F=(f_{1},\ldots,f_{n})\in\mathcal{F}. Suppose that n>1n>1, then

[F]S[F¯]S=\displaystyle[F]_{S}\star[\overline{F}]_{S}= [f1,,fn]S[fn¯,,f1¯]S=[f1,,fn,fn¯,,f1¯]S\displaystyle[f_{1},\ldots,f_{n}]_{S}\star[\overline{f_{n}},\ldots,\overline{f_{1}}]_{S}=[f_{1},\ldots,f_{n},\overline{f_{n}},\ldots,\overline{f_{1}}]_{S}
=\displaystyle= [f1,,fn1,fn1¯,,f1¯]S=[f1,f1¯]S,\displaystyle[f_{1},\dots,f_{n-1},\overline{f_{n-1}},\ldots,\overline{f_{1}}]_{S}=[f_{1},\overline{f_{1}}]_{S},
[F¯]S[F]S=\displaystyle[\overline{F}]_{S}\star[F]_{S}= [fn¯,,f1¯]S[f1,,fn]S=[fn¯,,f1¯,f1,,fn]S\displaystyle[\overline{f_{n}},\ldots,\overline{f_{1}}]_{S}\star[f_{1},\ldots,f_{n}]_{S}=[\overline{f_{n}},\ldots,\overline{f_{1}},f_{1},\ldots,f_{n}]_{S}
=\displaystyle= [fn¯,,f2¯,f2,,fn]S=[fn¯,fn]S,\displaystyle[\overline{f_{n}},\ldots,\overline{f_{2}},f_{2},\ldots,f_{n}]_{S}=[\overline{f_{n}},f_{n}]_{S},
[f1,f1¯]S[F]S=\displaystyle[f_{1},\overline{f_{1}}]_{S}\star[F]_{S}= [f1,f1¯]S[f1,,fn]S=[f1,f1¯,f1,fn]S\displaystyle[f_{1},\overline{f_{1}}]_{S}\star[f_{1},\ldots,f_{n}]_{S}=[f_{1},\overline{f_{1}},f_{1},\ldots f_{n}]_{S}
=\displaystyle= [f1,,fn]S=[F]S=[f1,,fn,fn¯,fn]S\displaystyle[f_{1},\ldots,f_{n}]_{S}=[F]_{S}=[f_{1},\ldots,f_{n},\overline{f_{n}},f_{n}]_{S}
=\displaystyle= [f1,,fn]S[fn¯,fn]S=[F]S[fn¯,fn]S\displaystyle[f_{1},\ldots,f_{n}]_{S}\star[\overline{f_{n}},f_{n}]_{S}=[F]_{S}\star[\overline{f_{n}},f_{n}]_{S}

If n=1n=1, we can find g,h:1Xg,h:\mathbb{Z}_{1}\rightarrow X such that their tails are eventually equal and F=(f1)=(g,h)F=(f_{1})=(g,h). Then we have that F¯=(f1¯)=(g,h)¯=(h¯,g¯)\overline{F}=(\overline{f_{1}})=\overline{(g,h)}=(\overline{h},\overline{g}). Therefore

[F]S[F¯]S=\displaystyle[F]_{S}\star[\overline{F}]_{S}= [g,h]S[h¯,g¯]S=[g,h,h¯,g¯]S=[f1,f1¯]S\displaystyle[g,h]_{S}\star[\overline{h},\overline{g}]_{S}=[g,h,\overline{h},\overline{g}]_{S}=[f_{1},\overline{f_{1}}]_{S}
[F¯]S[F]S=\displaystyle[\overline{F}]_{S}\star[F]_{S}= [h¯,g¯]S[g,h]S=[h¯,g¯,g,h]S=[f1¯,f1]S\displaystyle[\overline{h},\overline{g}]_{S}\star[g,h]_{S}=[\overline{h},\overline{g},g,h]_{S}=[\overline{f_{1}},f_{1}]_{S}
[f1,f1¯]S[F]S=\displaystyle[f_{1},\overline{f_{1}}]_{S}\star[F]_{S}= [g,h,h¯,g¯]S[g,h]S=[g,h,h¯,g¯,g,h]S=[g,h]S=[F]S\displaystyle[g,h,\overline{h},\overline{g}]_{S}\star[g,h]_{S}=[g,h,\overline{h},\overline{g},g,h]_{S}=[g,h]_{S}=[F]_{S}
=\displaystyle= [g,h,h¯,g¯,g,h]S=[g,h]S[h¯,g¯,g,h]S=[F]S[f1¯,f1]S.\displaystyle[g,h,\overline{h},\overline{g},g,h]_{S}=[g,h]_{S}\star[\overline{h},\overline{g},g,h]_{S}=[F]_{S}\star[\overline{f_{1}},f_{1}]_{S}.\qed
Lemma 4.19.

Let f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X bornologous maps such that there exist m,nm,n\in\mathbb{Z} such that f(nz)=g(mz)f(n-z)=g(m-z) for every zz\in\mathbb{Z}. Then, [f,f¯]S=[g,g¯]S[f,\overline{f}]_{S}=[g,\overline{g}]_{S}.

Proof.

Let f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X bornologous maps such that there exist m,nm,n\in\mathbb{Z} such that f(nz)=g(mz)f(n-z)=g(m-z) for every zz\in\mathbb{Z}.

Suppose that ff is eventually left periodic with T{1,2}T\in\{1,2\}, then gg satisfies that condition too. By 4.13, we obtain that [f,f¯]S=[h]S=[g,g¯][f,\overline{f}]_{S}=[h]_{S}=[g,\overline{g}] defining

h(z){f(z0) if z2f(z01) if z2+1\displaystyle h(z)\coloneqq\left\{\begin{array}[]{ll}f(z_{0})&\text{ if }z\in 2\mathbb{Z}\\ f(z_{0}-1)&\text{ if }z\in 2\mathbb{Z}+1\end{array}\right.

for z0z_{0} big enough.

Suppose that ff is not eventually left periodic with T{1,2}T\in\{1,2\}, then there exists n<nn^{\prime}<n such that f(n1)f(n+1)f(n^{\prime}-1)\neq f(n^{\prime}+1). Observe that

g(m(nn+1))=f(n1)f(n+1)=g(m(nn1)).\displaystyle g(m-(n-n^{\prime}+1))=f(n^{\prime}-1)\neq f(n^{\prime}+1)=g(m-(n-n^{\prime}-1)).

By 4.13, we have that [f]S=[h,f]S[f]_{S}=[h,f^{\prime}]_{S} and [g]S=[h,g]S[g]_{S}=[h,g^{\prime}]_{S} defining

h(z){f(n+z) if z0f(n) if z>0,\displaystyle h(z)\coloneqq\left\{\begin{array}[]{ll}f(n^{\prime}+z)&\text{ if }z\leq 0\\ f(n^{\prime})&\text{ if }z>0\end{array}\right., f(z){f(n) if z0f(n+z) if z>0,\displaystyle\ f^{\prime}(z)\coloneqq\left\{\begin{array}[]{ll}f(n^{\prime})&\text{ if }z\leq 0\\ f(n^{\prime}+z)&\text{ if }z>0\end{array}\right.,
g(z){f(n) if z0g(mn+n+z) if z>0.\displaystyle g^{\prime}(z)\coloneqq\left\{\begin{array}[]{ll}f(n^{\prime})&\text{ if }z\leq 0\\ g(m-n+n^{\prime}+z)&\text{ if }z>0.\end{array}\right.

Hence, [f,f¯]S=[h,f,f¯,h¯]S=[h,h¯]S=[h,g,g¯,h¯]S=[g,g¯]S[f,\overline{f}]_{S}=[h,f^{\prime},\overline{f^{\prime}},\overline{h}]_{S}=[h,\overline{h}]_{S}=[h,g^{\prime},\overline{g^{\prime}},\overline{h}]_{S}=[g,\overline{g}]_{S}. ∎

Definition 4.20.

The extended groupoid of XX, π¯1(X)\overline{\pi}_{\leq 1}(X), is the groupoid with objects the symmetric maps to XX and morphisisms the strings.

Suppose that we have XX and YY semi-coarse spaces and a bornologous map h:XYh:X\rightarrow Y. If we have a string (f1,,fn)(f_{1},\ldots,f_{n}) on XX, then h(f1,,fn)(hf1,,hfn)h(f_{1},\ldots,f_{n})\coloneqq(hf_{1},\ldots,hf_{n}) is a string on YY. From this definition, we directly observe that the string are a coarse invariant.

Proposition 4.21.

Let X,YX,Y be isomorphic semi-coarse spaces. Then π¯1(X)π¯1(X)\overline{\pi}_{\leq 1}(X)\cong\overline{\pi}_{\leq 1}(X) as groupoids.

Along the following construction, we are going to define classes between affine functions. The behavior in the tails are much more important than the behavior in the middle of the functions, then we allow to delete some middle points whenever the results is still a bornologous map.

Definition 4.22.

Let f:1Xf:\mathbb{Z}_{1}\rightarrow X be a bornologous map. Let z0z_{0}\in\mathbb{Z} and x0Xx_{0}\in X.

  • We said that gg is the result of deleting z0z_{0} in ff, or we can delete z0z_{0} in ff, if

    g(z){f(z) if z<z0f(z+1) if zz0.\displaystyle g(z)\coloneqq\left\{\begin{array}[]{ll}f(z)&\text{ if }z<z_{0}\\ f(z+1)&\text{ if }z\geq z_{0}.\end{array}\right.

    is bornologous. This is denoted as fd(z0)gf\xrightarrow{d(z_{0})}g.

  • We said that gg is the result of adding x0x_{0} at z0z_{0} in ff, or we can add x0x_{0} at z0z_{0}, if

    g(z){f(z) if z<z0x0 if z=z0f(z1) if z>z0.\displaystyle g(z)\coloneqq\left\{\begin{array}[]{ll}f(z)&\text{ if }z<z_{0}\\ x_{0}&\text{ if }z=z_{0}\\ f(z-1)&\text{ if }z>z_{0}.\end{array}\right.

    is bornologous. This is denoted as fa(z0,x0)gf\xrightarrow{a(z_{0},x_{0})}g.

Lemma 4.23.

Let XX be a semi-coarse space and f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X be bornologous maps. Then, fd(z0)gf\xrightarrow{d(z_{0})}g if and only if ga(z0,f(z0))fg\xrightarrow{a(z_{0},f(z_{0}))}f.

Proof.

Suppose that fd(z0)gf\xrightarrow{d(z_{0})}g. By definition, we have that g(z)=f(z)g(z)=f(z) for every z<z0z<z_{0} and g(z)=f(z+1)g(z)=f(z+1) for every zz0z\geq z_{0}. Moreover, the result of adding f(z0)f(z_{0}) at z0z_{0} in gg is

h(z)={g(z) if z<z0f(z0) if z=z0g(z1) if z>z0.\displaystyle h(z)=\left\{\begin{array}[]{ll}g(z)&\text{ if }z<z_{0}\\ f(z_{0})&\text{ if }z=z_{0}\\ g(z-1)&\text{ if }z>z_{0}.\end{array}\right.

Replacing the values of gg, we have that

h(z)={f(z) if z<z0f(z0) if z=z0f(z1+1)=f(z) if z>z0.\displaystyle h(z)=\left\{\begin{array}[]{ll}f(z)&\text{ if }z<z_{0}\\ f(z_{0})&\text{ if }z=z_{0}\\ f(z-1+1)=f(z)&\text{ if }z>z_{0}.\end{array}\right.

Thus ga(z0,f(z0))fg\xrightarrow{a(z_{0},f(z_{0}))}f.

Suppose that ga(z0,f(z0))fg\xrightarrow{a(z_{0},f(z_{0}))}f. By definition we have that f(z)=g(z)f(z)=g(z) for every z<z0z<z_{0} and f(z)=g(z1)f(z)=g(z-1) for every z>z0z>z_{0}. Moreover, the result of deleting z0z_{0} in ff is

h(z)={f(z) if z<z0f(z+1) if zz0.\displaystyle h(z)=\left\{\begin{array}[]{ll}f(z)&\text{ if }z<z_{0}\\ f(z+1)&\text{ if }z\geq z_{0}.\end{array}\right.

Replacing the values of ff, we have that

h(z)={g(z) if z<z0g(z+11)=g(z) if zz0.\displaystyle h(z)=\left\{\begin{array}[]{ll}g(z)&\text{ if }z<z_{0}\\ g(z+1-1)=g(z)&\text{ if }z\geq z_{0}.\end{array}\right.

Thus fa(z0)ff\xrightarrow{a(z_{0})}f. ∎

Lemma 4.24.

Let XX be a semi-coarse space and f:1Xf:\mathbb{Z}_{1}\rightarrow X be a bornologous map, then

fa(z0){f(z) if z<z0f(z0) if z=z0f(z1) if z>z0\displaystyle f_{a(z_{0})}\coloneqq\left\{\begin{array}[]{ll}f(z)&\text{ if }z<z_{0}\\ f(z_{0})&\text{ if }z=z_{0}\\ f(z-1)&\text{ if }z>z_{0}\end{array}\right.

is a bornologous map, and therefore fa(z0,f(z0))fa(z0)f\xrightarrow{a(z_{0},f(z_{0}))}f_{a(z_{0})}.

Proof.

Let f:1Xf:\mathbb{Z}_{1}\rightarrow X be a bornologous maps. We take the subsets X1(,z0]X_{1}\coloneqq(-\infty,z_{0}], X2[z0,z0+1]X_{2}\coloneqq[z_{0},z_{0}+1] and X3[z0+1,)X_{3}\coloneqq[z_{0}+1,\infty). Observe that every controlled set in 1\mathbb{Z}_{1} is an union of controlled sets in 1X1\mathbb{Z}_{1}\mid_{X_{1}}, 1X2\mathbb{Z}_{1}\mid_{X_{2}} and 1X3\mathbb{Z}_{1}\mid_{X_{3}}, and X=X1X2X3X=X_{1}\cup X_{2}\cup X_{3}. Thus, we can use 2.3 to prove that fa(z0)f_{a(z_{0})} is a bornologous map. Observe that fa(z0)X1=fX1f_{a(z_{0})}\mid_{X_{1}}=f\mid_{X_{1}} and fa(z0)X3=f[z0,)f_{a(z_{0})}\mid_{X_{3}}=f\mid_{[z_{0},\infty)}, which are bornologous maps because they are restrictions of a bornologous map. Further, fa(z0)X2=f{z0}f_{a(z_{0})}\mid_{X_{2}}=f\mid_{\{z_{0}\}} which goes from just one point, then it is a bornologous map. Thus, fa(z0)f_{a(z_{0})} is a bornologous map. ∎

Definition 4.25.

Let XX be a semi-coarse space and f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X be bornologous maps. We say that fdgf\simeq_{d}g if there exist a finite sequence such that

f=f0α0f1α1αn1fn=g\displaystyle f=f_{0}\xrightarrow{\alpha_{0}}f_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}f_{n}=g

with αi{d(z)z}{a(z,x)z,xX}\alpha_{i}\in\{d(z)\mid z\in\mathbb{Z}\}\cup\{a(z,x)\mid z\in\mathbb{Z},x\in X\}.

Theorem 4.26.

d\simeq_{d} is an equivalence relation.

Proof.

Let XX be a semi-coarse space and f,g,h:1Xf,g,h:\mathbb{Z}_{1}\rightarrow X be bornologous maps.

Reflexive: Since ff is bornologous, then fa(z0)f_{a(z_{0})} is bornologous by 4.24. Moreover, by 4.23, we have that fa(z0)d(z0)ff_{a(z_{0})}\xrightarrow{d(z_{0})}f. Then

fa(z0,f(z0))fa(z0)d(z0)f\displaystyle f\xrightarrow{a(z_{0},f(z_{0}))}f_{a(z_{0})}\xrightarrow{d(z_{0})}f

Thus, fdgf\simeq_{d}g.

Symmetry: Suppose that fdgf\simeq_{d}g. Then there exist f1,,fn1:1Xf_{1},\ldots,f_{n-1}:\mathbb{Z}_{1}\rightarrow X bornologous maps and α0,,αn1{a(z,x)z,xX}{d(z)z}\alpha_{0},\ldots,\alpha_{n-1}\in\{a(z,x)\mid z\in\mathbb{Z},x\in X\}\cup\{d(z)\mid z\in\mathbb{Z}\} such that

f=f0α0f1α1αn1fn=g\displaystyle f=f_{0}\xrightarrow{\alpha_{0}}f_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}f_{n}=g

For every ii:

  • If αi=a(zi,xi)\alpha_{i}=a(z_{i},x_{i}), then define βid(zi)\beta_{i}\coloneqq d(z_{i}).

  • If αi=d(zi)\alpha_{i}=d(z_{i}), then define βia(zi,fi(zi))\beta_{i}\coloneqq a(z_{i},f_{i}(z_{i})).

Thus, applying 4.23, we have the sequence

g=fnβn1fn1βn2β0f0=f,\displaystyle g=f_{n}\xrightarrow{\beta_{n-1}}f_{n-1}\xrightarrow{\beta_{n-2}}\cdots\xrightarrow{\beta_{0}}f_{0}=f,

and gdfg\simeq_{d}f.

Transitive: Suppose that fdgf\simeq_{d}g and gdhg\simeq_{d}h.Then there exist f1,,fn1,g1,,gm1:1Xf_{1},\ldots,f_{n-1},g_{1},\ldots,g_{m-1}:\mathbb{Z}_{1}\rightarrow X bornologous maps and α0,,αn1,β1,,βm1{a(z,x)z,xX}{d(z)z}\alpha_{0},\ldots,\alpha_{n-1},\beta_{1},\ldots,\beta_{m-1}\in\{a(z,x)\mid z\in\mathbb{Z},x\in X\}\cup\{d(z)\mid z\in\mathbb{Z}\} such that

f=f0α0f1α1αn1fn=g,g=g0β0g1β1βm1fm=h.\displaystyle f=f_{0}\xrightarrow{\alpha_{0}}f_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}f_{n}=g,\ g=g_{0}\xrightarrow{\beta_{0}}g_{1}\xrightarrow{\beta_{1}}\cdots\xrightarrow{\beta_{m-1}}f_{m}=h.

We can observe that

f=f0α0f1α1αn1fn=g=g0β0g1β1βm1fm=h,\displaystyle f=f_{0}\xrightarrow{\alpha_{0}}f_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}f_{n}=g=g_{0}\xrightarrow{\beta_{0}}g_{1}\xrightarrow{\beta_{1}}\cdots\xrightarrow{\beta_{m-1}}f_{m}=h,

thus fdhf\simeq_{d}h. ∎

To work in the strings, we would like to see that this equivalence relation is well-defined in the affine classes. Suppose that fdgf\simeq_{d}g and take f^fA\hat{f}\in\langle f\rangle_{A}. By definition, there exists a sequence

ff0α0f1α1αn1fng,\displaystyle f\coloneqq f_{0}\xrightarrow{\alpha_{0}}f_{1}\xrightarrow{\alpha_{1}}\cdots\xrightarrow{\alpha_{n-1}}f_{n}\eqqcolon g,

and there exists mm\in\mathbb{Z} such that f^(z)=f(zm)\hat{f}(z)=f(z-m) for every zz\in\mathbb{Z}.

Define f^i:1X\hat{f}_{i}:\mathbb{Z}_{1}\rightarrow X with the relation f^i(z)=fi(zm)\hat{f}_{i}(z)=f_{i}(z-m) for every 0<in0<i\leq n.

First suppose that αi=d(zi)\alpha_{i}=d(z_{i}), then

fi+1(z)={fi(z) if z<zifi(z+1) if zzi=\displaystyle f_{i+1}(z)=\left\{\begin{array}[]{ll}f_{i}(z)&\text{ if }z<z_{i}\\ f_{i}(z+1)&\text{ if }z\geq z_{i}\end{array}\right.\ =\ f^i+1(z+m)={f^i(z+m) if z<zif^i(z+m+1) if zzi\displaystyle\hat{f}_{i+1}(z+m)=\left\{\begin{array}[]{ll}\hat{f}_{i}(z+m)&\text{ if }z<z_{i}\\ \hat{f}_{i}(z+m+1)&\text{ if }z\geq z_{i}\end{array}\right.
=\displaystyle=\ f^i+1(z)={f^i(z) if z<zi+mf^i(z+1) if zzi+m\displaystyle\hat{f}_{i+1}(z)=\left\{\begin{array}[]{ll}\hat{f}_{i}(z)&\text{ if }z<z_{i}+m\\ \hat{f}_{i}(z+1)&\text{ if }z\geq z_{i}+m\end{array}\right.

Then f^id(zi+m)f^i+1\hat{f}_{i}\xrightarrow{d(z_{i}+m)}\hat{f}_{i+1}.

On the other hand, suppose that αi=a(zi,xi)\alpha_{i}=a(z_{i},x_{i}), then

fi+1(z)={fi(z) if z<zixi if z=zifi(z+1) if z>zi=\displaystyle f_{i+1}(z)=\left\{\begin{array}[]{ll}f_{i}(z)&\text{ if }z<z_{i}\\ x_{i}&\text{ if }z=z_{i}\\ f_{i}(z+1)&\text{ if }z>z_{i}\end{array}\right.\ =\ f^i+1(z+m)={f^i(z+m) if z<zixi if z=zif^i(z+m+1) if zzi\displaystyle\hat{f}_{i+1}(z+m)=\left\{\begin{array}[]{ll}\hat{f}_{i}(z+m)&\text{ if }z<z_{i}\\ x_{i}&\text{ if }z=z_{i}\\ \hat{f}_{i}(z+m+1)&\text{ if }z\geq z_{i}\end{array}\right.
=\displaystyle=\ f^i+1(z)={f^i(z) if z<zi+mxi if z=zi+mf^i(z+1) if zzi+m\displaystyle\hat{f}_{i+1}(z)=\left\{\begin{array}[]{ll}\hat{f}_{i}(z)&\text{ if }z<z_{i}+m\\ x_{i}&\text{ if }z=z_{i}+m\\ \hat{f}_{i}(z+1)&\text{ if }z\geq z_{i}+m\end{array}\right.

Then f^ia(zi+m,xi)f^i+1\hat{f}_{i}\xrightarrow{a(z_{i}+m,x_{i})}\hat{f}_{i+1}. Therefore f^f^ng^\hat{f}\simeq\hat{f}_{n}\eqqcolon\hat{g}, and g^gA\hat{g}\in\langle g\rangle_{A}, what gives us the following definition.

Definition 4.27.

Let XX be a semi-coarse space and f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X be bornologous maps. We say that fAdgA\langle f\rangle_{A}\simeq_{d}\langle g\rangle_{A} if there exists g^gA\hat{g}\in\langle g\rangle_{A} such that fdg^f\simeq_{d}\hat{g}.

Observe that this relation is well defined for the procedure above; if we take f^f\hat{f}\in\langle f\rangle, there exists hg^A=gAh\in\langle\hat{g}\rangle_{A}=\langle g\rangle_{A} such that f^dh\hat{f}\simeq_{d}h.

Proposition 4.28.

Let XX be a semi-coarse space and f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X be eventually equal symmetric maps. Then fAdgA\langle f\rangle_{A}\simeq_{d}\langle g\rangle_{A}.

Proof.

Let f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X be eventually equal symmetric maps. Then there exists m,nm,n\in\mathbb{N} such that f(z+n)=g(z+m)f(z+n)=g(z+m) for every z0z\in\mathbb{N}_{0}

We first reduce ff and apply an analogous algorithm with gg. Since ff is symmetric, then f(n)=f(n)f(n)=f(-n) for every n0n\in\mathbb{N}_{0}; thus, applying 2.3, we obtain that the result of deleting 0 in ff, call it f1f_{1}, is a bornologous map. Moreover, the result of deleting 0 in f1f_{1}, call it f1f^{\prime}_{1} is bornologous because f1=(f1)a(1)f_{1}=(f_{1}^{\prime})_{a(-1)}. Observe that, if we define f^1:1X\hat{f}_{1}:\mathbb{Z}_{1}\rightarrow X such that f^1(z)=f1(z1)\hat{f}_{1}(z)=f_{1}^{\prime}(z-1), we obtain the following

f^1(z)={f(z+1)z0f(z1)z<0, and fd(0)f1d(0)f1f^1A.\displaystyle\hat{f}_{1}(z)=\left\{\begin{array}[]{ll}f(z+1)&z\geq 0\\ f(z-1)&z<0,\end{array}\right.\text{ and }f\xrightarrow{d(0)}f_{1}\xrightarrow{d(0)}f^{\prime}_{1}\in\langle\hat{f}_{1}\rangle_{A}.

We can repeat this process from 11 to n1n-1; it means, we call fi+1f_{i+1} the result of deleting 0 in f^i\hat{f}_{i}, fi+1f_{i+1}^{\prime} the result of deleting 0 in fi+1f_{i+1}, and we define f^i+1:1X\hat{f}_{i+1}:\mathbb{Z}_{1}\rightarrow X such that f^i+1(z)=fi+1(z1)\hat{f}_{i+1}(z)=f^{\prime}_{i+1}(z-1), obtaining that

f^i+1(z)={f^i(z+1)z0f^i(z1)z<0, and f^id(0)fi+1d(0)fi+1f^i+1A.\displaystyle\hat{f}_{i+1}(z)=\left\{\begin{array}[]{ll}\hat{f}_{i}(z+1)&z\geq 0\\ \hat{f}_{i}(z-1)&z<0,\end{array}\right.\text{ and }\hat{f}_{i}\xrightarrow{d(0)}f_{i+1}\xrightarrow{d(0)}f^{\prime}_{i+1}\in\langle\hat{f}_{i+1}\rangle_{A}.

Therefore, fAdf^nA\langle f\rangle_{A}\simeq_{d}\langle\hat{f}_{n}\rangle_{A}. Using the same algorithm with gg until m1m-1 we obtain that gAdg^mA\langle g\rangle_{A}\simeq_{d}\langle\hat{g}_{m}\rangle_{A}. It only leaves to verify that, by definition

f^n={f(z+n)z0f(zn)z<0=\displaystyle\hat{f}_{n}=\left\{\begin{array}[]{ll}f(z+n)&z\geq 0\\ f(z-n)&z<0\end{array}\right.\ =\ f^n={g(z+m)z0g(zm)z<0\displaystyle\hat{f}_{n}=\left\{\begin{array}[]{ll}g(z+m)&z\geq 0\\ g(z-m)&z<0\end{array}\right.
=\displaystyle\ =\ g^m\displaystyle\hat{g}_{m}

Concluding that fAdgA\langle f\rangle_{A}\simeq_{d}\langle g\rangle_{A}. ∎

In the result above we prove that 4.25 allows to make homotopy deformations in the path which joins two tails. This proposition is the reason for the fundamental groupoid (4.32) contains the semi-coarse fundamental group for every point in the space.

Proposition 4.29.

Let XX be a semi-coarse space and f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X such that there exists NN\in\mathbb{N} satisfying

  • f(,N][N,)=g(,N][N,)f\mid_{(-\infty,-N]\cup[N,\infty)}=g\mid_{(-\infty,-N]\cup[N,\infty)}.

  • f[N,N]scg[N,N]f\mid_{[-N,N]}\simeq_{sc}g\mid_{[-N,N]} through H:[N,N]×1XH:[-N,N]\times\mathbb{Z}_{1}\rightarrow X such that H(N,)=f(N)H(-N,-)=f(-N) and H(N,)=f(N)H(N,-)=f(N).

Then fAdgA\langle f\rangle_{A}\simeq_{d}\langle g\rangle_{A}.

Proof.

Let XX be a semi-coarse space and f,g:1Xf,g:\mathbb{Z}_{1}\rightarrow X and {0,1}\{0,1\} with the subspace structure. First we consider a homotopy with just one step. Suppose that there exists NN\in\mathbb{N} satisfying

  • f(,N][N,)=g(,N][N,)f\mid_{(-\infty,-N]\cup[N,\infty)}=g\mid_{(-\infty,-N]\cup[N,\infty)}.

  • There exists H:[N,N]×{0,1}XH:[-N,N]\times\{0,1\}\rightarrow X bornologous map such that H(N,)=f(N)H(-N,-)=f(-N), H(N,)=f(N)H(N,-)=f(N), H(z,0)=f(z)H(z,0)=f(z) and H(z,1)=g(z)H(z,1)=g(z).

By the condition two, observe that {(f(i),g(i+1)),(g(i+1),f(i+2))}\{(f(i),g(i+1)),(g(i+1),f(i+2))\} is controlled in XX for every ii\in\mathbb{Z}, in particular for NiN-N\leq i\leq N. Thus, we obtain the followin

fa(N+1)f1a(N+1,g(N+1)f1′′d(N+2)f1′′′.\displaystyle f\xrightarrow{a(-N+1)}f^{\prime}_{1}\xrightarrow{a(-N+1,g(-N+1)}f^{\prime\prime}_{1}\xrightarrow{d(-N+2)}f^{\prime\prime\prime}_{1}.

Observe that f(z)=f1′′′(z)f(z)=f^{\prime\prime\prime}_{1}(z) for every zN+1z\neq-N+1 and f1′′′(N+1)=g(N+1)f^{\prime\prime\prime}_{1}(-N+1)=g(-N+1). We can repeat this interaction as

fi′′′a(N+i+1)fi+1a(N+i+1,g(N+i+1)fi+1′′d(N+i+2)fi+1′′′,\displaystyle f^{\prime\prime\prime}_{i}\xrightarrow{a(-N+i+1)}f^{\prime}_{i+1}\xrightarrow{a(-N+i+1,g(-N+i+1)}f^{\prime\prime}_{i+1}\xrightarrow{d(-N+i+2)}f^{\prime\prime\prime}_{i+1},

obtaining that fi′′′(z)=fi+1′′′(z)f_{i}^{\prime\prime\prime}(z)=f^{\prime\prime\prime}_{i+1}(z) for every zN+i+1z\neq-N+i+1 and fi+1′′′(N+i+1)=g(N+i+1)f^{\prime\prime\prime}_{i+1}(-N+i+1)=g(-N+i+1). Making this algorithm until i=2N2i=2N-2 we obtain f2N1′′′=gf^{\prime\prime\prime}_{2N-1}=g. Thus fdgf\simeq_{d}g.

In the general case, we have that there exists NN\in\mathbb{N} satisfying

  • f(,N][N,)=g(,N][N,)f\mid_{(-\infty,-N]\cup[N,\infty)}=g\mid_{(-\infty,-N]\cup[N,\infty)}.

  • f[N,N]scg[N,N]f\mid_{[-N,N]}\simeq_{sc}g\mid_{[-N,N]} through H:[N,N]×1XH:[-N,N]\times\mathbb{Z}_{1}\rightarrow X such that H(N,)=f(N)H(-N,-)=f(-N) and H(N,)=f(N)H(N,-)=f(N).

Since HH is a homotopy, there exists a<ba<b\in\mathbb{Z} such that H(z,x)=f(x)H(z,x)=f(x) for every xax\leq a and H(z,x)=g(z)H(z,x)=g(z) for every xbx\geq b. Define fi:[N,N]Xf_{i}:[-N,N]\rightarrow X such that fi(z)=H(z,i)f_{i}(z)=H(z,i) for every aiba\leq i\leq b Hence, taking HiH[N,N]×{i,i+1}H_{i}\coloneqq H\mid_{[-N,N]\times\{i,i+1\}} we obtain that

f=fadfa+1ddfb1dfb=g\displaystyle f=f_{a}\simeq_{d}f_{a+1}\simeq_{d}\cdots\simeq_{d}f_{b-1}\simeq_{d}f_{b}=g

Concluding that fdgf\simeq_{d}g. ∎

Definition 4.30.

Let (f1,,fn)(f_{1},\ldots,f_{n}) and (g1,,gn)(g_{1},\ldots,g_{n}) be nn strings. We say that (f1,,fn)d(g1,,gn)(f_{1},\ldots,f_{n})\simeq_{d}(g_{1},\ldots,g_{n}) if fidgif_{i}\simeq_{d}g_{i}.

Definition 4.31.

Let F,GF,G\in\mathcal{F}. We say that we can convert and deform FF in GG if there exists a finite sequence of arrows

FF0α0F1α1αn1FnG\displaystyle F\coloneqq F_{0}\simeq_{\alpha_{0}}F_{1}\simeq_{\alpha_{1}}\cdots\simeq_{\alpha_{n-1}}F_{n}\eqqcolon G

with αi{S,d}\alpha_{i}\in\{S,d\}. We express this relation as FGF\simeq G.

Definition 4.32.

The fundamental groupoid of XX, π1(X)\pi_{\leq 1}(X), is the groupoid with objects the symmetric maps to XX and morphisisms the classes of strings under the delete and deform relation.

5. Relative Fundamental Groupoid and Van Kampen Theorem

In the semi-coarse fundamental groupoid, Van Kampen Theorem is not true in general. We can restrict the tails of the maps from 1\mathbb{Z}_{1} to get a relative version of the fundamental groupoid; there we obtain the theorem with the correct partition of the space.

Definition 5.1.

Let XX be a semi-coarse space and 𝒰2X\mathcal{U}\subset 2^{X}. The 𝒰\mathcal{U}-relative fundamental groupoid, π1(X,𝒰)\pi_{\leq 1}(X,\mathcal{U}), is the category with the following ingredients:

  • Ob(π1(X,𝒰))Ob(\pi_{\leq 1}(X,\mathcal{U})) are the symmetric maps which are eventually equal to a symmetric map with image contained in some U𝒰U\in\mathcal{U}.

  • Homπ1(X,𝒰)(f,g)Hom_{\pi_{\leq 1}(X,\mathcal{U})}(f,g) are strings F=[f1,,fn]F=[f_{1},\ldots,f_{n}] such that for every i{0,,n}i\in\{0,\ldots,n\} there exists Ui𝒰U_{i}\in\mathcal{U} such that

    • The left tail of fif_{i} and the right tail if fi+1f_{i+1} land in UiU_{i} for i1,,n1i\in{1,\ldots,n-1}.

    • The left tail of f1f_{1} lands in U0U_{0}.

    • The right tail of fnf_{n} lans in UnU_{n}.

First, we introduce a lemma which mimics the implementation of the lebesgue covering lemma [Brown_2006] in the van Kampen Theorem for the semi-coarse fundamental groupoid.

Lemma 5.2.

Let XX be a semi-coarse space, A,BXA,B\subset X well-split XX and f:1Xf:\mathbb{Z}_{1}\rightarrow X. Additionally, suppose that each tail is within AA or BB. Then, there exists f1,,fn:1Xf_{1},\ldots,f_{n}:\mathbb{Z}_{1}\rightarrow X such that [f]=[f1][f2][fn][f]=[f_{1}]\star[f_{2}]\star\ldots\star[f_{n}], fi()Af_{i}(\mathbb{Z})\subset A or fi()Bf_{i}(\mathbb{Z})\subset B for every ii and the right tail of fif_{i} is the same as the left tail of fi+1f_{i+1}

Proof.

Let XX be a semi-coarse space, A,BXA,B\subset X well-split XX and f:1Xf:\mathbb{Z}_{1}\rightarrow X. Trying to simplify the redaction, we say that the left tail is eventually in a set UU and the right tail is eventually in a set VV.

If f()Af(\mathbb{Z})\subset A or f()Bf(\mathbb{Z})\subset B, we already have what we wanted to. Suppose that f()Af(\mathbb{Z})\not\subset A and f()Bf(\mathbb{Z})\not\subset B. Without loss of generality, suppose that UAU\subset A. There exists zz\in\mathbb{Z} such that f(z)BAf(z)\in B-A. Then we take the set z1<z2<<zkz_{1}<z_{2}<\ldots<z_{k} such that f(zi)BAf(z_{i})\in B-A and f(zi+1)ABf(z_{i+1})\in A-B for every i{1,3,,2k2+1}i\in\{1,3,\ldots,2\lfloor\frac{k}{2}\rfloor+1\}.

UULeft tailAAf(z1)f(z_{1})BBf(z2)f(z_{2})VVRight tail
Figure 5. Dividing ff

If f(zi)f(z_{i}) belongs to ABA\cap B, we don’t do anything. Otherwise, since A,BA,B split well XX, there exists λi:{0,1,2}X\lambda_{i}:\{0,1,2\}\rightarrow X in AABBA\sqcup_{A\cap B}B such that λi(0)=f(zi1)\lambda_{i}(0)=f(z_{i}-1), λi(2)=f(zi)\lambda_{i}(2)=f(z_{i}) and λi(1)AB\lambda_{i}(1)\in A\cap B. Then ff belongs to the same class of ff^{\prime}, with the addition of λi(1)\lambda_{i}(1) among zi1z_{i}-1 and ziz_{i}. Hence we obtain a new sequence z1<z2<<zkz^{\prime}_{1}<z^{\prime}_{2}<\ldots<z^{\prime}_{k} such that f(zi)ABf(z_{i}^{\prime})\in A\cap B, and f(zi)BAf(z^{\prime}_{i})\in B-A and f(zi+1)ABf(z^{\prime}_{i+1})\in A-B for every i{1,3,,2k2+1}i\in\{1,3,\ldots,2\lfloor\frac{k}{2}\rfloor+1\}.

UULeft tailAAf(z1)f^{\prime}(z^{\prime}_{1})BBf(z2)f^{\prime}(z_{2})VVRight tail
Figure 6. Dividing ff^{\prime}

We define the following maps for i{1,k1}i\in\{1,\ldots k-1\}

fL(z){f(z) if zz1f(z1) if z>z1,\displaystyle f_{L}(z)\coloneqq\left\{\begin{array}[]{ll}f^{\prime}(z)&\text{ if }z\leq z^{\prime}_{1}\\ f^{\prime}(z^{\prime}_{1})&\text{ if }z>z^{\prime}_{1},\end{array}\right. fR(z){f(zk) if z<zkf(z) if zzk,\displaystyle f_{R}(z)\coloneqq\left\{\begin{array}[]{ll}f^{\prime}(z^{\prime}_{k})&\text{ if }z<z^{\prime}_{k}\\ f^{\prime}(z)&\text{ if }z\geq z^{\prime}_{k},\end{array}\right.
fi(z){f(zi) if z<zif(z) if zizzi+1f(zi+1) if z>zi+1.\displaystyle f_{i}(z)\coloneqq\left\{\begin{array}[]{ll}f^{\prime}(z^{\prime}_{i})&\text{ if }z<z^{\prime}_{i}\\ f^{\prime}(z)&\text{ if }z^{\prime}_{i}\leq z\leq z^{\prime}_{i+1}\\ f^{\prime}(z^{\prime}_{i+1})&\text{ if }z>z^{\prime}_{i+1}.\end{array}\right.

Hence, by construction fL()Af_{L}(\mathbb{Z})\subset A, fi()Bf_{i}(\mathbb{Z})\subset B and fi+1()Af_{i+1}(\mathbb{Z})\subset A for i{1,3,}i\in\{1,3,\ldots\}, and fRf_{R} is subset of AA or BB. In addition, the right tail of fif_{i} and the left tail of fi+1f_{i+1} are the same and constant. ∎

Previous to the main theorem’s proof, we are going to consider a connected semi-coarse space; remembering that this implies that ABA\cap B\neq\varnothing if A,BA,B well-split XX, for 3.8.

Theorem 5.3.

Let XX be a connected semi-coarse space and 𝒰2X\mathcal{U}\subset 2^{X}such that U𝒰U\in\mathcal{U} is connected. Consider A,BXA,B\subset X such that

  1. (1)

    A,BA,B well-split XX,

  2. (2)

    for every component of AA, BB or ABA\cap B there exists UU such that meets the component, and

  3. (3)

    for every U𝒰U\in\mathcal{U}, UAU\subset A and UBU\subset B.

Then the diagram

is a pushout.

Proof.

Let XX be a connected semi-coarse space and 𝒰2X\mathcal{U}\subset 2^{X}such that U𝒰U\in\mathcal{U} is connected.

Let GG be a groupoid and the following diagram be commutative

We proceed to construct a map h:π1(X,𝒰)Gh:\pi_{\leq 1}(X,\mathcal{U})\rightarrow G such that is the unique such that j2=hi2j_{2}=hi_{2} and j1=hi1j_{1}=hi_{1}.

First we work with the objects; we define h:Ob(π1(X,𝒰))Ob(G)h:Ob(\pi_{\leq 1}(X,\mathcal{U}))\rightarrow Ob(G) such that h(fA)=j1(fA)h(\langle f\rangle_{A})=j_{1}(\langle f\rangle_{A}) if fAOb(π1(A,𝒰A))\langle f\rangle_{A}\in Ob(\pi_{\leq 1}(A,\mathcal{U}\cap A)) and h(fA)=j2(fA)h(\langle f\rangle_{A})=j_{2}(\langle f\rangle_{A}) if fAOb(π1(B,𝒰B))\langle f\rangle_{A}\in Ob(\pi_{\leq 1}(B,\mathcal{U}\cap B)). Since every U𝒰U\in\mathcal{U} satisfies that UAU\subset A or UBU\subset B, then for every fAOb(π1(X,𝒰))\langle f\rangle_{A}\in Ob(\pi_{\leq 1}(X,\mathcal{U})) the image of gg is completely contained in some U𝒰U\in\mathcal{U}, then it is totally contained in AA or BB. Hence fAOb(π1(A,𝒰A))\langle f\rangle_{A}\in Ob(\pi_{\leq 1}(A,\mathcal{U}\cap A)) or fAOb(π1(B,𝒰B))\langle f\rangle_{A}\in Ob(\pi_{\leq 1}(B,\mathcal{U}\cap B)). If gA\langle g\rangle_{A} is object of both categories, then is well-defined because j2iB=j1iAj_{2}i_{B}=j_{1}i_{A}.

By 5.2, we can write

[f]=[f1,L][f1,1][f1,k1][f1,R,f2,L][f2,1][fn,R]\displaystyle[f]=[f_{1,L}]\star[f^{\prime}_{1,1}]\star\ldots[f^{\prime}_{1,k_{1}}]\star[f_{1,R},f_{2,L}]\star[f^{\prime}_{2,1}]\star\ldots\star[f_{n,R}]

such that every factor is a morphism in π1(A,𝒰A)\pi_{\leq 1}(A,\mathcal{U}\cap A) or in π1(B,𝒰B)\pi_{\leq 1}(B,\mathcal{U}\cap B). Observe that every factor have constant tails, except the left tail of f1,Lf_{1,L} and the right tail of fn,Rf_{n,R}. In addition, this constant tails are in a component ABA\cap B, then, there exists Ui,j𝒰U_{i,j}\in\mathcal{U} which meets such component. Then, there exists a bornologous map γi,j:{0,mi,j}X\gamma^{\prime}_{i,j}:\{0,m_{i,j}\}\rightarrow X totally contained in the component such that γi,j(0)\gamma^{\prime}_{i,j}(0) is the right tail of fi,jf_{i,j} and fi,j(mi,j)Ui,jf_{i,j}(m_{i,j})\in U_{i,j}. Through γi,j\gamma^{\prime}_{i,j}, we define

γi,j(z){γi,j(0) if z0γi,j(z) if 0<z<mi,jγi,j(mi,j) if zmi,j.\displaystyle\gamma_{i,j}(z)\coloneqq\left\{\begin{array}[]{ll}\gamma^{\prime}_{i,j}(0)&\text{ if }z\leq 0\\ \gamma^{\prime}_{i,j}(z)&\text{ if }0<z<m_{i,j}\\ \gamma^{\prime}_{i,j}(m_{i,j})&\text{ if }z\leq m_{i,j}.\end{array}\right.

with fi,0=fi,Lf_{i,0}=f_{i,L} and fi,kn+1=fi,Rf_{i,k_{n}+1}=f_{i,R}. Therefore we take

[gi,j]=[γi,j1][fi,j][γi,j¯] and [gi,R,gi+1,L]=[γi,kn][fi,j][γi+1,0¯]\displaystyle[g_{i,j}]=[\gamma_{i,j-1}]\star[f_{i,j}]\star[\overline{\gamma_{i,j}}]\text{ and }[g_{i,R},g_{i+1,L}]=[\gamma_{i,k_{n}}]\star[f_{i,j}]\star[\overline{\gamma_{i+1,0}}]

with the exceptions [g1,L]=[f1,L][γ1,0][g_{1,L}]=[f_{1,L}]\star[\gamma_{1,0}] and [gn,R]=[γn,kn¯][fn,R][g_{n,R}]=[\overline{\gamma_{n,k_{n}}}]\star[f_{n,R}]. Every gi,jg_{i,j} is completly contained in AA or BB, then [gi,j][g_{i,j}] is a string in AA or in BB. Thus

[f]=ι1,0[g1,L]ι1,1[g1,1]ιn,kn+1[gn,R]\displaystyle[f]=\iota_{1,0}[g_{1,L}]\star\iota_{1,1}[g_{1,1}]\star\ldots\iota_{n,k_{n}+1}[g_{n,R}]

where ιi,j{i1,i2}\iota_{i,j}\in\{i_{1},i_{2}\}. In this order of ideas, we can suspect that, in case such hh homomorphism exists, then

h[f]=κ1,0[g1,L]κ1,1[g1,1]κn,kn+1[gn,R]\displaystyle h[f]=\kappa_{1,0}[g_{1,L}]\star\kappa_{1,1}[g_{1,1}]\star\ldots\kappa_{n,k_{n}+1}[g_{n,R}]

such that κi,j=j1\kappa_{i,j}=j_{1} if ιi,j=i1\iota_{i,j}=i_{1} and κi,j=j2\kappa_{i,j}=j_{2} if ιi,j=i2\iota_{i,j}=i_{2}.

Hence this assure the uniqueness of hh, and proves that π1(X,𝒰)\pi_{\leq 1}(X,\mathcal{U}) is generates as a grupoid by the images of π1(A,𝒰A)\pi_{\leq 1}(A,\mathcal{U}\cap A) and π1(B,𝒰U)\pi_{\leq 1}(B,\mathcal{U}\cap U).

The laborious part of this proof is to show that this construction does not depend of the election of the element of [f][f] or the election of γi,j\gamma^{\prime}_{i,j}. To prove that it doesn’t depend of the election of the γi,j\gamma^{\prime}_{i,j}, we take Ui,jU^{\prime}_{i,j} which meets the component and γi,j,2:{0,,mi,j,2}X\gamma^{\prime}_{i,j,2}:\{0,\ldots,m_{i,j,2}\}\rightarrow X totally contained in the component such that γi,j,2(0)\gamma^{\prime}_{i,j,2}(0) is the right tail of fi,jf_{i,j} and fi,j(mi,j,2)Ui,jf_{i,j}(m_{i,j,2})\in U^{\prime}_{i,j}. We define γi,j,2\gamma_{i,j,2} as the same way we define γi,j\gamma_{i,j} and observe that:

κi,j[gi,j]κi>,j>[gi>,j>]=\displaystyle\kappa_{i,j}[g_{i,j}]\kappa_{i^{>},j^{>}}[g_{i^{>},j^{>}}]= κi,j[γi<,j<¯,fi,j,γi,j,γi,j¯,γi,j,2,γi,j,2¯,γi,j]κi>,j>[γi,j¯,fi>,j>γi>,j>]\displaystyle\kappa_{i,j}[\overline{\gamma_{i^{<},j^{<}}},f_{i,j},\gamma_{i,j},\overline{\gamma_{i,j}},\gamma_{i,j,2},\overline{\gamma_{i,j,2}},\gamma_{i,j}]\kappa_{i^{>},j^{>}}[\overline{\gamma_{i,j}},f_{i^{>},j^{>}}\gamma_{i^{>},j^{>}}]
=\displaystyle= κi,j[γi<,j<¯,fi,j,γi,j,2,γi,j,2¯,γi,j]κi>,j>[γi,j¯,fi>,j>γi>,j>]\displaystyle\kappa_{i,j}[\overline{\gamma_{i^{<},j^{<}}},f_{i,j},\gamma_{i,j,2},\overline{\gamma_{i,j,2}},\gamma_{i,j}]\kappa_{i^{>},j^{>}}[\overline{\gamma_{i,j}},f_{i^{>},j^{>}}\gamma_{i^{>},j^{>}}]
=\displaystyle= κi,j[γi<,j<¯,fi,j,γi,j,2]κi>,j>[γi,j,2¯,γi,j,γi,j¯,fi>,j>γi>,j>]\displaystyle\kappa_{i,j}[\overline{\gamma_{i^{<},j^{<}}},f_{i,j},\gamma_{i,j,2}]\kappa_{i^{>},j^{>}}[\overline{\gamma_{i,j,2}},\gamma_{i,j},\overline{\gamma_{i,j}},f_{i^{>},j^{>}}\gamma_{i^{>},j^{>}}]
=\displaystyle= κi,j[γi<,j<¯,fi,j,γi,j,2]κi>,j>[γi,j,2¯,fi>,j>γi>,j>].\displaystyle\kappa_{i,j}[\overline{\gamma_{i^{<},j^{<}}},f_{i,j},\gamma_{i,j,2}]\kappa_{i^{>},j^{>}}[\overline{\gamma_{i,j,2}},f_{i^{>},j^{>}}\gamma_{i^{>},j^{>}}].

With a similar argument, we observe that the image neither depends of the election of the value of f(zk)f^{\prime}(z^{\prime}_{k}) in the construction 5.2. Consider that other λk:{0,1,2}X\lambda^{\prime}_{k}:\{0,1,2\}\rightarrow X in AABBA\sqcup_{A\cap B}B such that λk(0)=f(zk1)\lambda^{\prime}_{k}(0)=f(z_{k}-1) and λk(2)=fzk\lambda^{\prime}_{k}(2)=f_{z_{k}}. Since A,BA,B well-split XX, there exists a path α:{0,,m}AB\alpha:\{0,\ldots,m^{\prime}\}\rightarrow A\cap B such that α(0)=f(zk)\alpha(0)=f^{\prime}(z^{\prime}_{k}) and α(m)=λk(1)\alpha(m^{\prime})=\lambda^{\prime}_{k}(1). We define γi,j,2=αγi,j\gamma_{i,j,2}=\alpha\star\gamma_{i,j}, fi,j′′f^{\prime\prime}_{i,j} equal to fi,jf_{i,j} but replacing its right tail by λk(1)\lambda^{\prime}_{k}(1) and fi>,j>′′f^{\prime\prime}_{i^{>},j^{>}} is equal to fi>,j>f_{i^{>},j^{>}} but replacing its left tail by λk(1)\lambda^{\prime}_{k}(1). Thus

κi,j[gi,j]κi>,j>[gi>,j>]=\displaystyle\kappa_{i,j}[g_{i,j}]\kappa_{i^{>},j^{>}}[g_{i^{>},j^{>}}]= κi,j[γi<,j<¯,fi,j,γi,j]κi>,j>[γi,j¯,fi>,j>,γi>,j>]\displaystyle\kappa_{i,j}[\overline{\gamma_{i^{<},j^{<}}},f_{i,j},\gamma_{i,j}]\kappa_{i^{>},j^{>}}[\overline{\gamma_{i,j}},f_{i^{>},j^{>}},\gamma_{i^{>},j^{>}}]
=\displaystyle= κi,j[γi<,j<¯,fi,j′′,γi,j,2]κi>,j>[γi,j,2¯,fi>,j>′′,γi>,j>]\displaystyle\kappa_{i,j}[\overline{\gamma_{i^{<},j^{<}}},f^{\prime\prime}_{i,j},\gamma_{i,j,2}]\kappa_{i^{>},j^{>}}[\overline{\gamma_{i,j,2}},f^{\prime\prime}_{i^{>},j^{>}},\gamma_{i^{>},j^{>}}]
=\displaystyle= κi,j[γi<,j<¯,fi,j′′,γi,j,2]κi>,j>[γi,j,2¯,fi>,j>′′,γi>,j>]\displaystyle\kappa_{i,j}[\overline{\gamma_{i^{<},j^{<}}},f^{\prime\prime}_{i,j},\gamma_{i,j,2}]\kappa_{i^{>},j^{>}}[\overline{\gamma_{i,j,2}},f^{\prime\prime}_{i^{>},j^{>}},\gamma_{i^{>},j^{>}}]

The last equality follow from we always form the graph in Figure 7.

α(i)\alpha(i)α(i+1)\alpha(i+1)λk(0)\lambda^{\prime}_{k}(0)λk(2)\lambda^{\prime}_{k}(2)
Figure 7. Deleting α\alpha’s

Now, to see that it does not depend of the representative, it is enough to work with the arrows defined in 4.7, 4.13 and 4.22. Then, we define F(f1,,fn)F\coloneqq(f_{1},\ldots,f_{n}) and check case by case:

Fdop(i)G)F\xrightarrow{d_{op}(i)}G) We can apply the same algorithm as above to fif_{i}, getting that

[fi]=[gi,L,gi,1,,gi,ki,gi,R],\displaystyle[f_{i}]=[g_{i,L},g_{i,1},\ldots,g_{i,k_{i}},g_{i,R}],

and then we have obtain that

[fi¯]=[gi,R¯,gi,ki¯,gi,1¯,gi,L¯],\displaystyle[\overline{f_{i}}]=[\overline{g_{i,R}},\overline{g_{i,k_{i}}},\ldots\overline{g_{i,1}},\overline{g_{i,L}}],

Thus,

[F]=[f1,,fi1,gi,L,gi,L¯,,fn]=[G].\displaystyle[F]=[f_{1},\ldots,f_{i-1},g_{i,L},\overline{g_{i,L}},\ldots,f_{n}]=[G].

Faop(i,g)G)F\xrightarrow{a_{op}(i,g)}G) The same as the case above, but applying the procedure to gg.

Fm(i)G)F\xrightarrow{m(i)}G) We merge maps in four different ways. We work in the whole case up the second one, because is totally analogous to the first one

  • fif_{i} is periodic with T{1,2,3}T\in\{1,2,3\} and fi+1f_{i+1} is not periodic. Suppose that the right tail of fif_{i} is in U𝒰U\in\mathcal{U} and, without loss of generality, UAU\subset A. Then fi()UAf_{i}(\mathbb{Z})\subset U\subset A. Thus j1([fi])j_{1}([f_{i}]) is defined. Divide fi+1f_{i+1} as we construct in 5.2

    [fi+1]=[gi+1,L,gi+1,1,,gi+1,ni,gi+1,R]\displaystyle[f_{i+1}]=[g_{i+1,L},g_{i+1,1},\ldots,g_{i+1,n_{i}},g_{i+1,R}]

    Hence gi+1,L()Ag_{i+1,L}(\mathbb{Z})\subset A and [fi,gi+1,L]=[gL′′]π1(A,𝒰A)[f_{i},g_{i+1,L}]=[g^{\prime\prime}_{L}]\in\pi_{\leq 1}(A,\mathcal{U}\cap A), with gL′′g^{\prime\prime}_{L} the merging of fi,gi+1,Lf_{i},g_{i+1,L} (4.13). Thus

    j1[gi1,R,fi,gi+1,L]=\displaystyle j_{1}[g_{i-1,R},f_{i},g_{i+1,L}]= j1[gi1,R,gi+1,L′′]\displaystyle j_{1}[g_{i-1,R},g^{\prime\prime}_{i+1,L}]

    It only rest to observe that the merging of fi,fi+1f_{i},f_{i+1}, g′′g^{\prime\prime} is divided as

    [g′′]=[gL′′,fi+1,1,,fi+1,R].\displaystyle[g^{\prime\prime}]=[g^{\prime\prime}_{L},f_{i+1,1},\ldots,f_{i+1,R}].

    Getting that hh is well defined for this way to merge.

  • fi,fi+1f_{i},f_{i+1} are not periodic, and fif_{i} is eventually right periodic with T{1,2,3}T\in\{1,2,3\} or is not eventually right periodic. We already know that there exists a,ba^{\prime},b^{\prime}\in\mathbb{Z} such that fi(a+z)=fi+1(bz)f_{i}(a^{\prime}+z)=f_{i+1}(b^{\prime}-z) for every z0z\geq 0 and f(a1)fi+1(b+1)f(a^{\prime}-1)\neq f_{i+1}(b^{\prime}+1). Let’s call gg the merging of fif_{i} and fi+1f_{i+1} (4.13). By 5.2, we have

    [fi]=\displaystyle[f_{i}]= [fi,L,fi,1,,fi,ni,fi,R],\displaystyle[f_{i,L},f_{i,1},\ldots,f_{i,n_{i}},f_{i,R}],
    [fi+1]=\displaystyle[f_{i+1}]= [fi+1,L,fi+1,1,,fi+1,ni,fi+1,R],\displaystyle[f_{i+1,L},f_{i+1,1},\ldots,f_{i+1,n_{i}},f_{i+1,R}],
    [g′′]=\displaystyle[g^{\prime\prime}]= [gi,L′′,gi,1′′,,gi,n′′,gi,R′′].\displaystyle[g^{\prime\prime}_{i,L},g^{\prime\prime}_{i,1},\ldots,g^{\prime\prime}_{i,n^{\prime}},g^{\prime\prime}_{i,R}].

    Observe that aa^{\prime} lands in one fi,jf_{i,j} with j{0,,ni+1}j\in\{0,\ldots,n_{i}+1\} and fi+1,jf_{i+1,j^{\prime}} with j{0,,ni+1+1}j^{\prime}\in\{0,\ldots,n_{i+1}+1\}. Then we have that:

    [fi,fi+1]=[fi,L,fi,1,,fi,j,fi+1,j,fi+1,j+1,fi+1,R]=[g′′]\displaystyle[f_{i},f_{i+1}]=[f_{i,L},f_{i,1},\ldots,f_{i,j},f_{i+1,j^{\prime}},f_{i+1,j^{\prime}+1},f_{i+1,R}]=[g^{\prime\prime}]
  • fifi+1¯Af_{i}\in\langle\overline{f_{i+1}}\rangle_{A} and fif_{i} is eventually left periodic with T{1,2}T\in\{1,2\}. The procedure is the same as the previous case.

Fa(i,f,g)G)F\xrightarrow{a(i,f,g)}G) The sketchs of the proofs are the same as merging. You separate the maps through 5.2 and see how they work together.

In the cases of d\simeq_{d}, we can only work with specific maps; thus, we take fif_{i} and gg

fid(j)g)f_{i}\xrightarrow{d(j)}g) Consider the the zkz_{k} in fif_{i} as in 5.2. If zk<j<zk+1z_{k}<j<z_{k+1}, there is no problem in the construction when we remove the element in jj. Consider that j=zkj=z_{k} for some kk.

Without loss of generality, suppose that fi(zk)Bf_{i}(z_{k})\in B, and observe that {fi(j1),fi(j+1))}𝒱\{f_{i}(j-1),f_{i}(j+1))\}\in\mathcal{V}, since fid(j)gf_{i}\xrightarrow{d(j)}g. If fi(j1)ABf_{i}(j-1)\in A\cap B, there is nothing to do, because we select zkz_{k} as zkz^{\prime}_{k}. Assume that fi(j1)ABf_{i}(j-1)\notin A\cap B. Then we look for the first element between zkz_{k} and zk+1z_{k+1} such that is in BAB-A, let’s call this elements z^k\hat{z}_{k}.

Suppose that z^k=zk+1\hat{z}_{k}=z_{k}+1, then {(fi(zk1),fi(zk+1))}AABB\{(f_{i}(z_{k}-1),f_{i}(z_{k}+1))\}\notin A\sqcup_{A\cap B}B. Hence, there exists λi,k,λi,k:{0,1,2}X\lambda_{i,k},\lambda^{\prime}_{i,k}:\{0,1,2\}\rightarrow X in AABBA\sqcup_{A\cap B}B such that λi,k(0)=λi,k(0)=fi(zk1)\lambda_{i,k}(0)=\lambda^{\prime}_{i,k}(0)=f_{i}(z_{k}-1), λi,k(1)=λi,k(1)\lambda_{i,k}(1)=\lambda^{\prime}_{i,k}(1), λi,k(2)=fi(zk)\lambda_{i,k}(2)=f_{i}(z_{k}), and λi,k(2)=fi(zk+1)\lambda^{\prime}_{i,k}(2)=f_{i}(z_{k}+1).

fi(zk)f_{i}(z_{k})fi(zk+1)f_{i}(z_{k}+1)fi(zk1)f_{i}(z_{k}-1)λi,k(1)\lambda_{i,k}(1)
Figure 8. Deleting a point

Since A,BA,B well-split XX, then we can select f(zk)f^{\prime}(z^{\prime}_{k}) of the original partition as λi,k(1)\lambda_{i,k}(1) because there is a path between them. Thus, the sequence λi,k(1),f(zk+1),,f(zk+11)\lambda_{i,k}(1),f(z_{k}+1),\ldots,f(z_{k+1}-1) is equivalent in π1(B,𝒰B)\pi_{\leq 1}(B,\mathcal{U}\cap B) to λi,k(1),f(zk),f(zk+1),f(zk+11)\lambda_{i,k}(1),f(z_{k}),f(z_{k}+1),\ldots f(z_{k+1}-1). Getting what we wanted.

1-10111-1111-1λi,k(1)\lambda_{i,k}(1)111-1λi,k(1)\lambda_{i,k}(1)011
Figure 9. Representation of the process; fi(zk+1+1)ABf_{i}(z_{k+1}+1)\notin A\cap B and zk=0z_{k}=0 and z^k=1\hat{z}_{k}=1

fia(j,x0)g)f_{i}\xrightarrow{a(j,x_{0})}g) The case is analogous to the previous one. We only observe that happens if we add a new point which interrupts the sequence of element on AA or BB and we add one adequate point to recover the last case. ∎

Theorem 5.4.

Let XX be a semi-coarse space and 𝒰2X\mathcal{U}\subset 2^{X}, and write 𝒰\mathcal{U}^{\prime} the collection of all of the components of the elements of 𝒰\mathcal{U}. Consider A,BXA,B\subset X such that

  1. (1)

    A,BA,B well-split XX,

  2. (2)

    for every component of AA, BB ABA\cap B within a component of XX which meets Uin𝒰U\cup_{U\\ in\mathcal{U}}U, there exists U𝒰U^{\prime}\in\mathcal{U}^{\prime} such that meets the component, and

  3. (3)

    for every U𝒰U^{\prime}\in\mathcal{U}^{\prime}, UAU^{\prime}\subset A and UBU^{\prime}\subset B.

Then the diagram

is a pushout.

Proof.

We can divide XX in all its components and apply XαX_{\alpha} and apply 5.3 to every XαX_{\alpha} and Xα𝒰X_{\alpha}\cap\mathcal{U}^{\prime}. Obtaining the result. ∎

We conclude the section with two consequences of this theorem.

Corollary 5.5.

Let XX be a semi-coarse space and 𝒰={U}\mathcal{U}=\{U\} for some UXU\subset X , and write 𝒰\mathcal{U}^{\prime} the collection of all of the components of UU. Consider A,BXA,B\subset X such that

  1. (1)

    A,BA,B well-split XX,

  2. (2)

    for every component of AA, BB ABA\cap B within a component of XX which meets UU, UU meets the component, and

  3. (3)

    for every U𝒰U^{\prime}\in\mathcal{U}^{\prime}, UAU^{\prime}\subset A and UBU^{\prime}\subset B.

Then the diagram

is a pushout.

Corollary 5.6.

Let XX be a semi-coarse space and 𝒰={{x}xX}\mathcal{U}=\{\{x\}\mid x\in X\}. Consider A,BXA,B\subset X such that A,BA,B well-split XX. Then the diagram

is a pushout. In addition, the groupoid π1(X,𝒰)\pi_{\leq 1}(X,\mathcal{U}) contains all of the π1sc(X,x)\pi_{1}^{sc}(X,x).

6. Future Work

Most of the work in this document was to develop the fundamental groupoid and introduce the van Kampen theorem in this category. In the future, we would like to explore what other semi-coarse invariants we can study in an almost locally way with well-split notion. We did not find this concept in the literature and we hope this gives us some light about other constructions in this category.

Other authors have found relations between the coarse homotopy groups and the topological homotopy groups in compact manifolds (see [Weighill_LiftingCoarseHom_2021, Mitchener_2020].) We would like to study topological coarse spaces or topological semi-coarse spaces through this new approximation. Although we need deeper work, we expect that we will be capable to study open manifolds together its coarse structure (or maybe the semi coarse one.) In the same sense, we would like to observe the behavior of semi-coarse spaces adding other structure coming from “topological constructs”, for example closure spaces or limit spaces.

Appendix A Coarse Homotopy Fails in Semi-Coarse

Many coarse invariants has been studied before, and a natural question is why don’t we adapt some of them and study semi-coarse spaces. The problems with this kind of approximations are diverse, but generally: in some cases we do not recover the same information for coarse spaces, in others we lose some properties. In this brief appendix, we mention an example, the coarse homotopy [Mitchener_2020], and look at some of the problems with using semi-coarse spaces. We would like to remind that the semi-coarse homotopy groups are trivial in coarse spaces (3.2.19, [rieser2023semicoarse]).

In this section, +\mathbb{R}_{+} is endowed with the coarse structure

A{+×+r>0 such that d(x,y)r(x,y)A}.\displaystyle A\subset\{\mathbb{R}_{+}\times\mathbb{R}_{+}\mid\exists r>0\text{ such that }d(x,y)\leq r\ \forall(x,y)\in A\}.

In addition, we remind that in a coarse spaces:

Definition A.1.

Let (X,𝒱)(X,\mathcal{V}) and (Y,𝒲)(Y,\mathcal{W}) be coarse spaces

  • BXB\subset X is bounded in (X,𝒱)(X,\mathcal{V}) if there exists xXx\in X and A𝒱A\in\mathcal{V} such that

    A[x]{yX(x,y)A}=B.\displaystyle A[x]\coloneqq\{y\in X\mid(x,y)\in A\}=B.
  • A map f:XYf:X\rightarrow Y is proper if f1(B)f^{-1}(B) is bounded in (X,𝒱)(X,\mathcal{V}) for every BB bounded in (Y,𝒲)(Y,\mathcal{W}).

  • A map f:XYf:X\rightarrow Y bornologous and proper is called coarse map.

Definition A.2 (pp-Cylinder; [Mitchener_2020], 2.1).

Let XX be a coarse space, and let p:X+p:X\rightarrow\mathbb{R}_{+} be a coarse map. Then we define the pp-cylinder

IpX{(x,t)X×+tp(x)+1}\displaystyle I_{p}X\coloneqq\{(x,t)\in X\times\mathbb{R}_{+}\mid t\leq p(x)+1\}

We have inclusions i0:XIpXi_{0}:X\rightarrow I_{p}X and i1:XIpXi_{1}:X\rightarrow I_{p}X defined by the formulas i0(x)=(x,0)i_{0}(x)=(x,0) and i(x)=(x,p(x)+1)i(x)=(x,p(x)+1), respectively. We also have the canonical projection q:IpXXq:I_{p}X\rightarrow X defined by the formula q(x,t)=xq(x,t)=x.

Definition A.3 (Coarse Homotopy; [Mitchener_2020], 2.2).

Let XX and YY be coarse spaces. A coarse homotopy is a coarse map H:IpXYH:I_{p}X\rightarrow Y for some coarse map p:X+p:X\rightarrow\mathbb{R}_{+}.

We call coarse maps f0:XYf_{0}:X\rightarrow Y and f1:XYf_{1}:X\rightarrow Y coarsely homotopic if there is a coarse homotopy H:IpXYH:I_{p}X\rightarrow Y such that f0=Hi0f_{0}=H\circ i_{0} and f1=H1i1f_{1}=H_{1}\circ i_{1}.

This map HH is termed a coarse homotopy between the maps f0f_{0} and f1f_{1}.

We explore the following three modifications:

  1. (1)

    Let’s consider the definition of a bounded set AA as there exists a controlled set BB and xx such that A=B[x]A=B[x]. Then we might lose several coarse maps.

  2. (2)

    Consider just bornologous maps, then we might obtain a coarser equivalence relation which makes trivial some maps.

  3. (3)

    In the same case, for semi-coarse spaces, we lose transitive.

In every case, we provide an examples which illustrates what we claim in the last list.

Example A.3.1.

Let XX be a coarse space and consider f:1Xf:\mathbb{Z}_{1}\rightarrow X be a coarse map (a bornologous proper map.) Since (k,k+1)(k,k+1) is a controlled set in 1\mathbb{Z}_{1}, then we have that {f(0),f(5)}\{f(0),f(5)\} is a bounded set in XX, however {0,5}\{0,5\} is not in 1\mathbb{Z}_{1}. Thus, we don’t have any proper map from 1\mathbb{Z}_{1} to XX. (We can mimic this procedure replacing 1\mathbb{Z}_{1} for many semi-coarse spaces which are not coarse spaces.)

Example A.3.2.

Consider \mathbb{Z} with the coarse structure induced by the metric. We can observe that the map p:+p:\mathbb{Z}\rightarrow\mathbb{R}_{+} such that p(z)=|z|p(z)=|z| is bornologous (actually it is a coarse map.) Now we define H:Ip(X)XH:I_{p}(X)\rightarrow X such that H(x,t)=tH(x,t)=t; HH is a bornologous map and we can observe that it is not proper. Additionally, we observe that Hi0(x,t)=0H\circ i_{0}(x,t)=0 and Hi1(x,t)=|t|+1H\circ i_{1}(x,t)=|t|+1, obtaining that a map which goes to infinity is homotpic to a constant map.

Example A.3.3.

Consider XX as the semi-coarse space induced by the graph

0332211

and the bornologous map (which is not proper) p:1Xp:\mathbb{Z}_{1}\rightarrow X such that p(z)=0p(z)=0. Then we obtain that Ip(1)=1×[0,1]I_{p}(\mathbb{Z}_{1})=\mathbb{Z}_{1}\times[0,1]. Let’s define the following bornologous functions:

u:1X\displaystyle u:\mathbb{Z}_{1}\rightarrow X such that u(3n)=0,u(3n+1)=2,u(3n+2)=3 for every n\displaystyle\text{ such that }u(3n)=0,\ u(3n+1)=2,\ u(3n+2)=3\text{ for every }n\in\mathbb{Z}
m:1X\displaystyle m:\mathbb{Z}_{1}\rightarrow X such that u(3n)=0,u(3n+1)=u(3n+2)=2 for every n\displaystyle\text{ such that }u(3n)=0,\ u(3n+1)=u(3n+2)=2\text{ for every }n\in\mathbb{Z}
d:1X\displaystyle d:\mathbb{Z}_{1}\rightarrow X such that u(3n)=0,u(3n+1)=2,u(3n+2)=1 for every n\displaystyle\text{ such that }u(3n)=0,\ u(3n+1)=2,\ u(3n+2)=1\text{ for every }n\in\mathbb{Z}

Observe that we have the following homotopies:

H1(z,t)=u(z) if 0t<1,H1(z,t)=m(z) if t=1,\displaystyle H_{1}(z,t)=u(z)\text{ if }0\leq t<1,\ H_{1}(z,t)=m(z)\text{ if }t=1,
H2(z,t)=d(z) if 0t<1,H2(z,t)=m(z) if t=1\displaystyle H_{2}(z,t)=d(z)\text{ if }0\leq t<1,\ H_{2}(z,t)=m(z)\text{ if }t=1

However, there is no HH such that Hi0=uH\circ i_{0}=u and Hi1=dH\circ i_{1}=d because 1,3{1,3} is not controlled in XX.

As a last observation, we could try to replace +\mathbb{R}_{+} for a semi-coarse space in the definition of Ip(X)I_{p}(X) or in the codomain of HH. Observe that if (X,𝒱)(X,\mathcal{V}) is a connected coarse space, (Y,𝒲)(Y,\mathcal{W}) is a semi-coarse space and f:XYf:X\rightarrow Y is a bornologous map, then f(X)f(X) is a connected (in the coarse sense) space. This is a problem because it might limit our bornologous maps. To mention a brief example, if you replace +\mathbb{R}_{+} by +,r\mathbb{R}_{+,r} you will observe that every bornologous function pp maps a coarse space XX to p(X)p(X) with coarse structure 2p(X)×p(X)2^{p(X)\times p(X)}. Then we can find a semi-coarse space where the transitive fails.

Acknowledgments

The author would like to thanks to Antonio Rieser, who is the supervisor of my PhD project, for the several discussions about this paper and the possible directions this work could take, as well as his advice to include a brief discussion about the coarse homotopy groups. The author would also like to thanks Noe Barcenas for his feedback on an event at the Casa Mexicana de Matemáticas in Oaxaca and another event in the CCM in Morelia.

References