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

Homotopy Covers of Graphs

T. Chih and L. Scull
Abstract.

We develop a theory of ×\times-homotopy, fundamental groupoids and covering spaces that apply to non-simple graphs, generalizing existing results for simple graphs. We prove that ×\times-homotopies from finite graphs can be decomposed into moves which adjust at most one vertex at a time, generalizing the spider lemma of [CS1]. We define a notion of homotopy covering map and develop a theory of universal covers and deck transformations, generalizing [TardifWroncha, Matsushita] to non-simple graphs. We examine the case of reflexive graphs, where each vertex has at least one loop. We also prove that these homotopy covering maps satisfy a homotopy lifting property for arbitrary graph homomorphisms, generalizing path lifting results of [Matsushita, TardifWroncha].

1. Introduction

Covers of graphs were originally studied by viewing graphs as 1-dimensional topological spaces. From this perspective, many properties of basic topology have been extended to graphs, including the fundamental groupoid, covering spaces, universal covers and deck transformations [Angluin1, KwakNedela, Bass]. These results have been developed for non-simple graphs which allow loops and parallel edges.

Viewed as 1-dimensional topological spaces, the homotopies between graphs are limited. However, it is possible to develop theories of homotopies between graph homomorphisms that allow more interesting deformations and take into account more of the graph structure. There are two prominent theories of homotopies for graphs in the literature: AA-homotopy [Babson1, Barcelo1, BoxHomotopy, hardeman2019lifting] and ×\times-homotopy [CS1, CS2, Docht1, Docht2, CompGH, Kozlov1, Kozlov1, Kozlov3, KozlovShort]. Covers have been considered in the context of AA-homotopy in [hardeman2019lifting], where Hardeman shows that for graphs with no three or four cycles, graph covers lift AA-homotopies.

The goal of this paper is to understand what ×\times-homotopy, the fundamental group(oid) and covering maps of graphs look like in the context of graphs which allow loops and multiple parallel edges. We begin with some basic results about homotopies of graph homomorphisms between graphs with parallel edges in Section 2, showing that the spider lemma of [CS1] still holds, allowing us to decompose a ×\times-homotopy between finite graphs to a sequence of spider moves which adjust at most one vertex at a time. We then use this to define the fundamental groupoid of non-simple graphs in Section 3, following [CS2], and give a relationship between the fundamental groupoid of a reflexive graph and its unlooped counterpart in Propositon 3.8. In Section 4, we give a definition of homotopy covering maps for non-simple graphs and develop their properties, including a theory of deck transformations, for non-simple graphs. We again examine the case of reflexive graphs in particular. We conclude this section with proving that these homotopy covers satisfy the general homotopy lifting property for any graph homomorphism in Theorem 4.25. This last result generalizes results in the literature in two ways: in addition to extending to non-simple graphs, our result allows lifting of homotopies between any graph homomorphisms, and not just paths as in [Matsushita, TardifWroncha].

If we restrict to simple graphs, these results recover those in the literature in a few places. Our fundamental groupoid recovers that of Chih-Scull [CS2] when we allow loops but not parallel edges, and that of Sankar [Sankar] for simple graphs. Additionally, the spider lemma shows that the definition we use of ×\times-homotopy is equivalent to the definition of rr-homotopy in Matsushita [Matsushita] when r=2r=2, as this definiton allows homomorphisms to be shifted by rr edges (and hence r1r-1 vertices). For simple graphs, Matsushita develops covering space theory for rr-homotopy based on a neighbourhood complex. Lastly, for simple graphs Tardif-Wroncha [TardifWroncha] have a theory of \diamond-covers which lift any embedded 4-cycle, and in the appendix they give an alternate formulation which matches up with our Definition 4.4 of homotopy covers based on 2-neighbourhoods.

2. Graphs and Homotopy

We begin with the definition of our category of graphs, following [Bass, KwakNedela].

Definition 2.1.

The category of graphs 𝖦𝗉𝗁\mathsf{Gph} is defined by:

  • An object is a graph GG, consisting of a set of vertices V(G)={vλ}V(G)=\{v_{\lambda}\} and a set E(G)={eβ}E(G)=\{e_{\beta}\} of edges, with maps 0,1:E(G)V(G)\partial_{0},\partial_{1}:E(G)\to V(G) and an involution ee¯e\to\overline{e} of E(G)E(G) such that ie¯=1ie\partial_{i}\overline{e}=\partial_{1-i}e.

  • A homomorphism of graphs f:GHf:G\to H is given by a vertex map f:V(G)V(H)f:V(G)\to V(H) and an edge map f:E(G)E(H)f:E(G)\to E(H) such that f(ie)=if(e)f(\partial_{i}e)=\partial_{i}f(e) and f(e¯)=f(e)¯f(\overline{e})=\overline{f(e)}.

This category of graphs allows loops and multiple parallel edges between the same vertices. Bass [Bass] requires that the involution be fixed point free, while Kwak-Nedela [KwakNedela] allows ’semi-edges’ such that e¯=e\overline{e}=e.

In considering graphs which allow both loops and parallel edges, we will be comparing these to their single-edge and unlooped versions. Thus we will use the following notation.

Notation 2.2.

If GG is a graph, then:

  • GuG_{u} refers to the graph GG with all loops deleted: V(Gu)=V(G)V(G_{u})=V(G) and
    E(Gu)=E(G)\{e:0(d)=1(d)}E(G_{u})=E(G)\backslash\{e:\partial_{0}(d)=\partial_{1}(d)\}.

  • GsG_{s} refes to the single-edge graph with parallel edges have been identified: V(Gs)=V(G)V(G_{s})=V(G) and E(Gs)={[e]:eE(G) and [e]=[e] if 0(e)=0(e) and 1(e)=1(e)}E(G_{s})=\{[e]:e\in E(G)\textup{ and }[e]=[e^{\prime}]\textup{ if }\partial_{0}(e)=\partial_{0}(e^{\prime})\textup{ and }\partial_{1}(e)=\partial_{1}(e^{\prime})\}.

In this paper, we consider homotopy between graph homomorphisms described above, defined using the categorical product. This is sometimes referred to as ×\times-homotopy. This is the only version of homotopy that will appear in this paper, and we will refer to it simply as ‘homotopy’. It uses the following defintion.

Definition 2.3.

The categorical product G×HG\times H is defined by V(G×H)=V(G)×V(H)V(G\times H)=V(G)\times V(H) and E(G×H)=E(G)×E(H)E(G\times H)=E(G)\times E(H) with i(e1,e2)=(i(e1),i(e2))\partial_{i}(e_{1},e_{2})=(\partial_{i}(e_{1}),\partial_{i}(e_{2})) and (e1,e2)¯=(e¯1,e¯2)\overline{(e_{1},e_{2})}=(\overline{e}_{1},\overline{e}_{2}).

We define homotopy using the product with the interval graph InI_{n}.

Definition 2.4.

Let InI_{n} denote the looped path graph with vertices V(In)={0,1,2,,n}V(I_{n})=\{0,1,2,\dots,n\} and edges E(In)={i,ei,e¯i}E(I_{n})=\{\ell_{i},e_{i},\overline{e}_{i}\} for 1in1\leq i\leq n, where i(j)=j\partial_{i}(\ell_{j})=j and 0(ej)=j1,1(ej)=j\partial_{0}(e_{j})=j-1,\partial_{1}(e_{j})=j.

Observe that for any jj, the vertices of the form (v,j)(v,j) and edges of the form (e,j)(e,\ell_{j}) define a subgraph of G×InG\times I_{n} which is isomorphic to GG; we denote this subgraph G×{j}G\times\{j\}. The following definition is modeled on that of [Docht1].

Definition 2.5.

For f,g:GHf,g:G\to H, we say that ff is homotopic to gg, written fgf\simeq g, if there is a graph homomorphism Λ:G×InH\Lambda:G\times I_{n}\to H such that Λ|G×{0}=f\Lambda|_{G\times\{0\}}=f and Λ|G×{n}=g\Lambda|_{G\times\{n\}}=g. We write fgf\simeq g.

The introduction of multiple edges allows for multiple graph homomorphisms which agree on vertex sets. We show that these are all homotopic.

Lemma 2.6.

If f,g:GHf,g:G\to H are graph homomomorphisms which agree on vertices, then fgf\simeq g.

Proof.

We can define a homotopy Λ:G×I1H\Lambda:G\times I_{1}\to H by Λ(x,i)=f(x)=g(x)\Lambda(x,i)=f(x)=g(x) and

Λ(e,e)={f(e) if e=0g(e) otherwise {\Lambda}(e,e^{\prime})=\begin{cases}f(e)&\textup{ if }e^{\prime}=\ell_{0}\\ g(e)&\textup{ otherwise }\end{cases}

We can use this to show that GG is homotopy equivalent to the graph GsG_{s} where all parallel edges have been identified.

Proposition 2.7.

Let GG is a graph, and GsG_{s} the single-edge graph as in Notation 2.2. Then GsG_{s} is a strong deformation retract of GG.

Proof.

Define the projection π:GGs\pi:G\to G_{s} by the identity on vertices and the projection on edges. Define ι:GsG\iota:G_{s}\to G using the identity on vertices, and for each pair of edges [e],[e¯][e],[\overline{e}] choose an edge ee between the same vertices and define ι[e]=e\iota[e]=e and ι[e¯]=e¯\iota[\overline{e}]=\overline{e}. Then ιπ=idH\iota\pi=id_{H} and πιidG\pi\iota\simeq id_{G} by Lemma 2.6. ∎

Thus in any homotopy invariant construction, we may substitute GsG_{s} instead of GG.

When working with homotopies between graph homomorphisms that do not neccesarily agree on vertices, the following is a useful tool.

Proposition 2.8.

If GG is finite and f,g:GHf,g:G\to H then fgf\simeq g if and only if is a finite sequence of spider moves connecting ff and gg, where each spider move changes the value of ff by at most a single vertex, with the additional condition that if vertex vv has a loop, then any spider move that changes the value on vv must change to a connected vertex.

Proof.

The proof of [CS1], Proposition 4.4 can by adapted to the multi-edge case as follows. If ff and gg agree on vertices, then by Lemma 2.6 there is a one step homotopy between them. If f,gf,g differ by a single vertex vv then we can create a homotopy Λ:G×I1H\Lambda:G\times I_{1}\to H by Λ|G×{0}=f\Lambda|_{G\times\{0\}}=f and Λ|G×{1}=g\Lambda|_{G\times\{1\}}=g, and

Λ(e,e1)={f(e) if 0e=vg(e) if 1e=v{\Lambda}(e,e_{1})=\begin{cases}f(e)&\textup{ if }\partial_{0}e=v\\ g(e)&\textup{ if }\partial_{1}e=v\end{cases}

and

Λ(e,e¯1)={f(e) if 1e=vg(e) if 0e=v{\Lambda}(e,\overline{e}_{1})=\begin{cases}f(e)&\textup{ if }\partial_{1}e=v\\ g(e)&\textup{ if }\partial_{0}e=v\end{cases}

If vv has any loops i\ell_{i}, then we know there is an edge η\eta connecting f(v)f(v) to g(v)g(v) and we define Λ(i,e1)=η\Lambda(\ell_{i},e_{1})=\eta and Λ(i,e¯1)=η¯\Lambda(\ell_{i},\overline{e}_{1})=\overline{\eta}.

Conversely, suppose we have a homotopy fgf\simeq g; it is sufficient to consider a one-step homotopy Λ:G×I1H\Lambda:G\times I_{1}\to H between ff and gg and then induct to get longer homotopies. Suppose that GG has vertex set {v1,v2,,vn}\{v_{1},v_{2},\dots,v_{n}\}. We define a homotopy Λ~:G×InH\widetilde{\Lambda}:G\times I_{n}\to H on vertices by

Λ~(vi,k)={Λ(vi,0)=f(vi) if i>kΛ(vi,1)=g(vi) if ik\widetilde{\Lambda}(v_{i},k)=\begin{cases}\Lambda(v_{i},0)=f(v_{i})&\textup{ if }i>k\\ \Lambda(v_{i},1)=g(v_{i})&\textup{ if }i\leq k\end{cases}

and on edges by the following: if ee is an edge of GG with 0e=vi\partial_{0}e=v_{i} and 1e=vj\partial_{1}e=v_{j} then

Λ~(e,e)={Λ(e,0) if i>0(e),j>1(e)Λ(e,1) if i0(e),j1(e)Λ(e,e1) if i>0(e),j1(e)Λ(e,e¯1) if i0(e),j>1(e)\widetilde{\Lambda}(e,e^{\prime})=\begin{cases}\Lambda(e,\ell_{0})&\textup{ if }i>\partial_{0}(e^{\prime}),j>\partial_{1}(e^{\prime})\\ \Lambda(e,\ell_{1})&\textup{ if }i\leq\partial_{0}(e^{\prime}),j\leq\partial_{1}(e^{\prime})\\ \Lambda(e,e_{1})&\textup{ if }i>\partial_{0}(e^{\prime}),j\leq\partial_{1}(e^{\prime})\\ \Lambda(e,\overline{e}_{1})&\textup{ if }i\leq\partial_{0}(e^{\prime}),j>\partial_{1}(e^{\prime})\\ \end{cases}

This defines intermediate graph homomorphisms fk=Λ~(v,k)f_{k}=\widetilde{\Lambda}(v,k) in which fk,fk+1f_{k},f_{k+1} either agree on vertices or differ by a single vertex, and thus we get a sequence of spider moves joining ff to gg. ∎

3. The Fundamental Groupoid

In this section, we review the fundamental groupoid of [CS2] and extend it to the multi-edge case. We then consider the particular case of reflexive graphs.

Definition 3.1.

A walk of length nn is defined by a sequence of edges (e1e2en)(e_{1}e_{2}\dots e_{n}) such that 1(ei)=0(e1i)\partial_{1}(e_{i})=\partial_{0}(e_{1-i}). If 0(e1)=v\partial_{0}(e_{1})=v and 1(en)=w\partial_{1}(e_{n})=w we say that the walk starts at vv and ends at ww.

This is used to define what we refer to as the walk groupoid (called the fundamental groupoid in [KwakNedela, Bass]). We will present this in terms of the concept of ‘prunes’ from [CS2].

Definition 3.2.

When a walk α=(e1e2en)\alpha=(e_{1}e_{2}\dots e_{n}) satisfies e1i=e¯ie_{1-i}=\overline{e}_{i}, we refer to the walk α=(e1e2ei1ei+2en)\alpha^{\prime}=(e_{1}e_{2}\dots e_{i-1}e_{i+2}\dots e_{n}) that omits the pair eie1ie_{i}e_{1-i} as a prune of α\alpha, and α\alpha as an anti-prune of α\alpha^{\prime}.

Definition 3.3.

The walk groupoid 𝔚(G)\mathfrak{W}(G) is a groupoid with:

  • objects given by vertices vV(G)v\in V(G),

  • arrows starting vv and ending at ww given by prune classes of walks, where we identify any walks that are connected by a sequence of prunes and anti-prunes

  • composition defined by concatenation of walks.

This is presented in [Bass] as reduced walks, in which no prunes are possible. This walk groupoid does not identify homotopic walks. There are several equivalent definitions of a fundamental groupoid which does. We extend Definition 4.1 of [CS2] to the following.

Definition 3.4.

Let Π(G)\Pi(G) be the fundamental groupoid of GG defined by the following:

  • an object of Π(G)\Pi(G) is a vertex of the graph GG,

  • an arrow starting at v0v_{0} and ending at vnv_{n} in Π(G)\Pi(G) is given by a prune class of walks with the same start and end, up to homotopy rel endpoints,

  • composition of arrows is defined using concatenation of walks.

The results of Section 2 allow us to describe any homotopy of paths as a sequence of spider moves. Thus our fundamental groupoid consists of equivalence classes of walks, where two walks are equivalent if a sequence of prunes, anti-prunes and spider moves can transform one into the other.

Notation 3.5.

An arrow in Π(G)\Pi(G) is defined by a sequence of edges (e1e2en)(e_{1}e_{2}\dots e_{n}). However, Lemma 2.6 ensures that any choice of edges connecting the same sequence of vertices results in the same element of Π(G)\Pi(G). Thus in the rest of this section, we will notate arrows in the fundamental groupoid via their sequence of vertices (v0v1vn)(v_{0}v_{1}\dots v_{n}) and not specify edges.

We have seen that the addition of parallel edges does not change the homotopy behaviour, so will not be interesting when examining Π(G)\Pi(G). However, we are interested in the effect of including loops in our graphs.

Observation 3.6.

In the fundamental groupoid Π(G)\Pi(G)

  • Any loop =(vv)\ell=(vv) satisfies 2=(vvv)=1v\ell^{2}=(vvv)=1_{v} via a prune.

  • There is a spider move from (vvw)(vvw) to (vww)(vww).

  • If all vertices in a path starting at vv, and ending at ww are looped, we can always move the loops through the path and write the path as γ(ww)k\gamma*(ww)^{k} for γ\gamma containing no loops and k=0k=0 or 11.

In the remainder of this section, we examine the case of reflexive graphs, which we take here to mean that every vertex has at least one loop. Triangles play a particular role here, so we make the following definition.

Definition 3.7.

For a graph GG and its fundamental groupoid Π(G)\Pi(G), we define Π(G)/T\Pi(G)/T to be the quotient of the fundamental groupoid Π(G)\Pi(G) by the relation generated by triangles defined by an embedded C3C_{3}. Explicitly, we declare that if we have any embedded C3C_{3} defined by xyzxxyzx, then if α=α1α2\alpha=\alpha_{1}*\alpha_{2} in Π(G)\Pi(G) then α=α1(xyzx)α2\alpha=\alpha_{1}*(xyzx)*\alpha_{2} in Π(G)/T\Pi(G)/T. Thus we can insert or delete triangles from any path in Π(G)/T\Pi(G)/T.

Recall from Notation 2.2 that GuG_{u} refers to the unlooped graph obtained by removing all loops from GG.

Proposition 3.8.

Let GG be a reflexive graph. Then Π(G)Π(Gu)/T×/2\Pi(G)\cong\Pi(G_{u})/T\times\mathbb{Z}/2.

Proof.

The product groupoid Π(Gu)/T×/2\Pi(G_{u})/T\times\operatorname{\mathbb{Z}}/2 is defined as the product of categories, interpreting /2\operatorname{\mathbb{Z}}/2 as a category with a single object. Explicitly, Π(G)/T×/2\Pi(G)/T\times\operatorname{\mathbb{Z}}/2 has the same objects as Π(G)/T\Pi(G)/T, with arrows xyx\to y defined by (α,i)(\alpha,i) for α:xy\alpha:x\to y in Π(G)/T\Pi(G)/T and i=0,1/2i=0,1\in\operatorname{\mathbb{Z}}/2. The composition is defined by (α,i)(β,j)=(αβ,i+j (mod 2))(\alpha,i)\circ(\beta,j)=(\alpha\circ\beta,i+j\textup{ (mod 2)}).

We define the functor Ψ:Π(G)Π(Gu)/T×/2\Psi:\Pi(G)\to\Pi(G_{u})/T\times\mathbb{Z}/2 by Ψ(v)=v\Psi(v)=v on objects and Ψ(α)=(αu,i)\Psi(\alpha)=(\alpha_{u},i) on arrows, where αu\alpha_{u} denotes the walk α\alpha with all loops removed, and ii is the length of the walk α\alpha (mod 2).

To show that Ψ\Psi is well defined, we suppose that [α]=[α][\alpha]=[\alpha^{\prime}] in Π(G)\Pi(G). Then α\alpha and α\alpha^{\prime} are connected by a sequence of prunes, anti-prunes, and spider moves, all of which preserve the parity of the length. It is easy to see that any prune, anti-prune, or spider move not involving any loops has a corresponding move on the unlooped version, and that a prune involving a loop must just insert or delete a pair of loops, leaving the unlooped walk the same. So we consider the case when we have a spider move involving a loop. Unlooping (vvw)(vvw) and (vww)(vww) both result in (vw)(vw). If we replace α=uuv\alpha=uuv with α=uxv\alpha^{\prime}=uxv, then αu=uv\alpha_{u}=uv while αu=uxv=uxvuv=uv\alpha^{\prime}_{u}=uxv=uxvuv=uv since uxvuuxvu is a triangle. Hence the αu=αu\alpha_{u}=\alpha^{\prime}_{u} in Π(Gu)/T\Pi(G_{u})/T. Ψ\Psi respects composition since (αβ)u=αuβu(\alpha*\beta)_{u}=\alpha_{u}*\beta_{u}.

We define an inverse functor Φ:Π(Gu)/T×/2Π(G)\Phi:\Pi(G_{u})/T\times\mathbb{Z}/2\to\Pi(G) again using the identity on objects, and on arrows defining Φ(γ,i)=γ(ww)k\Phi(\gamma,i)=\gamma*(ww)^{k} where (ww)(ww) denotes the loop on the ending vertex ww of γ\gamma, and k=i+pk=i+p where pp is the mod 2 length of γ\gamma. This ensures that the parity of γ(ww)k\gamma*(ww)^{k} matches the parity of ii.

We check that this is well-defined: again, any prune, anti-prune, or spider move of γ\gamma gives a corresponding move on γ(ww)\gamma*(ww), so we just need to check what happens when we insert or remove a triangle. Suppose that we have two walks in GuG_{u} defined by γ\gamma and γ=γ1(xyzx)γ2\gamma^{\prime}=\gamma_{1}*(xyzx)*\gamma_{2} where γ1γ2=γ\gamma_{1}*\gamma_{2}=\gamma, so that we obtain γ\gamma^{\prime} by inserting a triangle into γ\gamma. Then the length of γ\gamma and γ\gamma^{\prime} have opposite parities, and so when we apply Φ\Phi to (γ,i)(\gamma,i) and to (γ,i)(\gamma^{\prime},i), one will be concatenated with a loop and the other will not. Assume that Φ(γ,i)=γ(ww)\Phi(\gamma,i)=\gamma*(ww) and Φ(γ,i)=γ\Phi(\gamma^{\prime},i)=\gamma^{\prime}. Then we use Observation 3.6 to get the following in Π(G)\Pi(G): γ(ww)=γ1(xyzx)γ2(ww)=γ1(xyzx)(xx)γ2=γ1(xyzxx)γ2=γ1(xyxxx)γ2=γ1(xyx)γ2=γ1γ2=γ\gamma^{\prime}*(ww)=\gamma_{1}*(xyzx)*\gamma_{2}*(ww)=\gamma_{1}*(xyzx)*(xx)*\gamma_{2}=\gamma_{1}*(xyzxx)*\gamma_{2}=\gamma_{1}*(xyxxx)*\gamma_{2}=\gamma_{1}*(xyx)*\gamma_{2}=\gamma_{1}*\gamma_{2}=\gamma. Therefore γ=γ(ww)(ww)=γ(ww)\gamma^{\prime}=\gamma^{\prime}*(ww)*(ww)=\gamma*(ww). A similar argument verifies well-definedness when Φ\Phi puts the loop on γ\gamma^{\prime}, and migrating loops through out paths ensures that Φ\Phi respects the concatenation operation.

Because all loops can be moved to the end of any walk in Π(G)\Pi(G), Ψ\Psi and Φ\Phi are inverse functors, giving the desired isomorphism of groupoids.

Corollary 3.9.

Suppose that GG is a reflexive graph with no embbedded C3C_{3}. Then the fundamental groupoid Π(G)Π(Gu)×/2\Pi(G)\simeq\Pi(G_{u})\times\operatorname{\mathbb{Z}}/2.

Example 3.10.

Consider the following graph GG and its unlooped subgraph GuG_{u}:

aabbccddeexxGG                                          aabbccddeexxGuG_{u}

The arrows of ΠGu\Pi G_{u} are walks in GuG_{u} up to edge shuffles, since these are the only possible spider moves. Therefore the closed walks (aexa),(abcdea)(aexa),(abcdea) do not commute in ΠGu\Pi G_{u}. When we add the loops and move to the graph GG, we introduce loops into our walks and the relations (aexa)=(aa),(vv)2=(v)(aexa)=(aa),(vv)^{2}=(v) and (uuv)=(uvv)(uuv)=(uvv). Thus we have commuting closed walks (abcde)(abcde) and (aexa)=(aa)(aexa)=(aa) and ΠGΠGu/(aexa)×/2\Pi G\cong\Pi G_{u}/\langle(aexa)\rangle\times\mathbb{Z}/2.

4. Homotopy Covers of Graphs

This section builds on the fundamental groupoids of Section 3 to develop the theory of homotopy covers of graphs. We define homotopy covering maps and show that they satisfy a version of the classical theory of universal covers and deck transformations, and describe the homotopy covers of reflexive graphs. We close by showing that these maps satisfy the homotopy lifting property for any graph homomorphisms in Theorem 4.25. Throughout this section, we will assume that all graphs are connected.

We begin with the classical definition of graph covers.

Definition 4.1.

[KwakNedela] A covering map is a graph morphism f:G~Gf:\widetilde{G}\to G such that

  • ff is an surjection on vertices and edges,

  • ff induces a bijection on neighbourhoods: let NE(v)N_{E}(v) denote the set of edges ee with 0(e)=v\partial_{0}(e)=v. Then for any vertex v~\tilde{v}, with f(v~)=vf(\tilde{v})=v, ff induces a bijection NE(v~)NE(f(v~))N_{E}(\tilde{v})\to N_{E}(f(\tilde{v})).

This reduces to the definition given by [Angluin1] for simple graphs, since in that case for e1,e2NE(v),1(e1)=1(e2)e_{1},e_{2}\in N_{E}(v),\partial_{1}(e_{1})=\partial_{1}(e_{2}) if and only if e1=e2e_{1}=e_{2}.

An important property of covering maps is the following path lifting result.

Lemma 4.2.

[Angluin1, Hatcher, KwakNedela] Suppose that f:G~Gf:\widetilde{G}\to G is a covering map. Then given a walk α\alpha in GG starting at vv defined by α=(e1e2en)\alpha=(e_{1}e_{2}\cdots e_{n}) and any v~f1(v)\tilde{v}\in f^{-1}(v), there exists a unique walk α~\tilde{\alpha} in G~\widetilde{G} starting at v~\tilde{v} defined by α~=(d1d2dn)\tilde{\alpha}=(d_{1}d_{2}\cdots d_{n}) such that f(α~)=αf(\tilde{\alpha})=\alpha.

The covers of Definition 4.1 do not incorporate the notion of homotopy of graphs. Thus we consider the following adaptation of the idea, which requires bijections on 2-neighbourhoods and not just neighbourhoods.

Definition 4.3.

For any vertex vV(G)v\in V(G) we define the extended neighborhood N2(v)N_{2}(v) to be the walks of length 22 which start at vv.

Our homotopy covering maps induce bijections on these 2-neighbourhoods.

Definition 4.4.

A surjective graph morphism between graphs f:G~Gf:\widetilde{G}\to G is a homotopy covering map if given any vV(G)v\in V(G) and v~V(G~)\tilde{v}\in V(\widetilde{G}) such that f~(v~)=v\tilde{f}(\tilde{v})=v, then ff induces a bijection N2(v~)N2(v)N_{2}(\tilde{v})\to N_{2}(v). Furthermore, this bijection respects endpoints in the sense that two walks in N2(v~),(d1d2),(d1d2)N_{2}(\tilde{v}),(d_{1}d_{2}),(d^{\prime}_{1}d^{\prime}_{2}) have 1(d2)=1(d2)\partial_{1}(d_{2})=\partial_{1}(d^{\prime}_{2}) in G~\widetilde{G} if and only if 1(f(d2))=1(f(d2))\partial_{1}(f(d_{2}))=\partial_{1}(f(d^{\prime}_{2})) in GG.

In working with these homotopy covering maps, we will find the following elementary observations extremely useful.

Observation 4.5.

Suppose that f:G~Gf:\widetilde{G}\to G is a homotopy covering map.

  • If a 2-walk in N2(v)N_{2}(v) is of the form ee¯e\overline{e}, then the lift must have the same form, since if we have a lift dddd^{\prime} then dd¯d\overline{d} defines a 2-walk which also projects down to ee¯e\overline{e} and so by uniqueness of lifting we must have d=d¯d^{\prime}=\overline{d}.

  • Any homotopy cover is a cover, since NE(v)N_{E}(v) is in bijection with 2-walks of the form ee¯e\overline{e}.

  • Parallel edges lift to parallel edges, since if e1,e2e_{1},e_{2} both have 0(ei)=v\partial_{0}(e_{i})=v and 1(ei)=w\partial_{1}(e_{i})=w then e1e¯2e_{1}\overline{e}_{2} is a 2-walk and so lifts to a 2-walk dddd^{\prime} from v~\widetilde{v} to v~\widetilde{v}^{\prime}. Then d¯\overline{d^{\prime}} is the unique lift of e2e_{2}.

We show that the classical theory of universal covers and deck transformations can be extended to homotopy covers. We will present these results in terms of the following groupoid notion.

Definition 4.6.

[Higgins] If xx is an object of a groupoid AA, the star of xx is the set of all arrows with source xx and is denoted AxA_{x}. Then f:ABf:A\to B is a covering of groupoids if ff induces a bijection on stars AxBf(x)A_{x}\to B_{f(x)} for all objects xx of AA.

When we apply this definition to the fundamental groupoid, we get another equivalent definition of homotopy covering map. We denote the star of Π(G)\Pi(G) at a vertex vv by Πv(G)\Pi_{v}(G).

Proposition 4.7.

A graph homomorphism f:G~Gf:\widetilde{G}\to G is a homotopy covering map if and only if ff is a cover as in Definiton 4.1 and the induced functor Π(f):Π(G~)Π(G)\Pi(f):\Pi(\widetilde{G})\to\Pi(G) is a covering of groupoids.

Proof.

Suppose that f:G~Gf:\widetilde{G}\to G is a homotopy covering map. To show that it induces a bijection on stars f:Πv~(G~)Πv(G)f_{*}:\Pi_{\widetilde{v}}(\widetilde{G})\to\Pi_{v}(G), we observe that since ff is a covering by Observation 4.5, Lemma 4.2 allows us to lift paths and so ff_{*} is surjective. To see that ff_{*} is also injective, we recall that equivalences in Π(G)\Pi(G) are generated by prunes, anti-prunes and spider moves. We again use Observation 4.5 to see that any prunable section of a walk ee¯e\overline{e} lifts to a similarly prunable section dd¯d\overline{d}, and that any spider move that changes edges but not vertices will lift to a similar edge move since parallel edges lift to parallel edges. Lastly, any spider move that adjusts one vertex will lift to a similar spider move because the bijection on second neighbourhoods N2(v)N_{2}(v) respects endpoints.

Conversely, suppose f:G~Gf:\widetilde{G}\to G is a graph cover which induces a covering of groupoids Π(f):Π(G~)Π(G)\Pi(f):\Pi(\widetilde{G})\to\Pi(G). We know that we can lift any 2-walk uniquely by Lemma 4.2. This lift must respect endpoints because if two 2-walks have the same endpoint, then they represent the same arrow in Πv(G)\Pi_{v}(G) and then the bijecton on stars ensures that their lifts represent the same arrow of Πv~(G~)\Pi_{\widetilde{v}}(\widetilde{G}) and hence have the same target vertex.

Definition 4.8.

Given a graph GG and vV(G)v\in V(G), we define a graph U=UvGU=U_{v}G and graph homomorphism ρ:UG\rho:U\to G as follows:

  • vertices of UvGU_{v}G are arrows in ΠvG\Pi_{v}G

  • edges between arrows α,β\alpha,\beta are defined by the edges between their endpoints: if α=(vv2w)\alpha=(vv_{2}\dots w) and β=(vv2w)\beta=(vv_{2}^{\prime}\dots w^{\prime}) then the edges dd with 0d=α\partial_{0}d=\alpha and 1d=β\partial_{1}d=\beta are given by the set {de:eE(G) such that 0(e)=w,1(e)=w}\{d_{e}:e\in E(G)\textup{ such that }\partial_{0}(e)=w^{\prime},\partial_{1}(e)=w\}.

  • the map ρ:UG\rho:U\to G is defined on vertices by the endpoint ρ(α)=ρ(vv2w)=w\rho(\alpha)=\rho(vv_{2}\dots w)=w and on edges by ρ(de)=e\rho(d_{e})=e.

Observation 4.9.

Recall from Notation 2.2 that GsG_{s} refers to the graph GG with parallel edges identified, and that Lemma 2.6 ensures that Π(G)Π(Gs)\Pi(G)\equiv\Pi(G_{s}). Thus UvGU_{v}G has the same vertices as UvGsU_{v}{G_{s}} but with appropriate parallel edges added above parallel edges of GG.

With this constrution, UvGU_{v}G is in fact a universal homotopy cover in the sense that it satisfies the appropriate universal properties.

Theorem 4.10.

Suppose that f:G~Gf:\widetilde{G}\to G is a homotopy cover and v~f1(v)\tilde{v}\in f^{-1}(v). Then there is a unique homotopy cover ρ~:UvGG~\tilde{\rho}:U_{v}G\to\widetilde{G} such that ρ=fρ~\rho=f\circ\tilde{\rho} and ρ~((v))=v~\tilde{\rho}((v))=\tilde{v}.

Sketch of Proof.

The proof of the above follows standard arugments as in [Angluin1, Bass, Hatcher]. Any walk α\alpha in GG lifts uniquely to a walk α~\widetilde{\alpha} starting at v~\tilde{v} in G~\widetilde{G}, and so we define ρ~(α)\tilde{\rho}(\alpha) to be the endpoint of this lift. This is easily extended to edges. ∎

The previous result implies that UvG,Uv~G~U_{v}G,U_{\tilde{v}}\widetilde{G} are covers of both G,G~G,\widetilde{G}, so they cover each other canonically and thus are isomorphic. Thus we also obtain the following result.

Corollary 4.11.

If G~G\widetilde{G}\to G is a homotopy cover, then the universal homotopy cover UU of GG is also the universal homotopy cover for G~\widetilde{G}.

We also have a theory of deck transformations of homotopy covering maps for non-simple graphs.

Definition 4.12.

Given a graph GG and the universal homotopy covering map ρ:UG\rho:U\to G, a deck transformation is an automorphsim fAut(U)f\in\operatorname{Aut}(U) such that ρf=ρ\rho\circ f=\rho. These form a subgroup of Aut(U)\operatorname{Aut}(U), which we denote by D(G)D(G).

We will show that this group is isomorphic to the isotropy group of the fundamental groupoid.

Definition 4.13.

Let ΠvvG\Pi_{v}^{v}G denote the isotropy group of ΠG\Pi G at vv, given by the arrows of ΠG\Pi_{G} starting and ending at vv. Explicitly, this is the group formed by walks that start and end at vv defined up to prunes and homotopy, under concatenation.

Because ΠG\Pi G is a connected groupoid, all the isotropy groups are isomorphic, and it does not matter which vertex vv we choose to look at. For γΠvvG\gamma\in\Pi_{v}^{v}G, we define a deck transformation as follows.

Definition 4.14.

Let γΠvv(G)\gamma\in\Pi_{v}^{v}(G). Define ψγ:UU\psi_{\gamma}:U\to U by αγα\alpha\mapsto\gamma*\alpha for vertices αV(U)\alpha\in V(U), and on edges by observing that the edges between α,β\alpha,\beta and γα,γβ\gamma*\alpha,\gamma*\beta are both in bijection with the edge set between the endpoints w,ww,w^{\prime} of α,β\alpha,\beta and so we define ψγ(de)=de\psi_{\gamma}(d_{e})=d^{\prime}_{e} for ee from ww to ww^{\prime}.

Each ψγ\psi_{\gamma} is deck transformation. In fact, every deck transformation is of this form, and thus we may identify the deck transformations of UU with Πvv(G)\Pi_{v}^{v}(G).

Theorem 4.15.

The map Φ:ΠvvGD(G)\Phi:\Pi_{v}^{v}G\to D(G) defined by γψγ\gamma\mapsto\psi_{\gamma} is an isomorphism of groups.

Proof.

The map Φ\Phi is a group homomorphism because Φ(γγ)(α)=γγα=ψγ(ψγ(α))=(Φ(γ)Φ(γ))(α)\Phi(\gamma*\gamma^{\prime})(\alpha)=\gamma*\gamma^{\prime}*\alpha=\psi_{\gamma}\left(\psi_{\gamma^{\prime}}(\alpha)\right)=\left(\Phi(\gamma)\circ\Phi(\gamma^{\prime})\right)(\alpha). To show that Φ\Phi is injective, suppose that γ,γΠvvG\gamma,\gamma^{\prime}\in\Pi_{v}^{v}G such that Φ(γ)=Φ(γ)\Phi(\gamma)=\Phi(\gamma^{\prime}). Then ψγ((v))=ψγ((v))\psi_{\gamma}((v))=\psi_{\gamma^{\prime}}((v)) and so γ=γ(v)=γ(v)=γ\gamma=\gamma*(v)=\gamma^{\prime}*(v)=\gamma^{\prime}.

To check that Φ\Phi is surjective, we let φD(G)\varphi\in D(G) and define γ=φ((v))\gamma=\varphi((v)). Since ρ(φ((v)))=ρ((v))=v\rho(\varphi((v)))=\rho((v))=v, the walk γ\gamma ends at vv and hence Φ(γ)=ψγ\Phi(\gamma)=\psi_{\gamma} is defined with ψγ((v))=φ((v))\psi_{\gamma}((v))=\varphi((v)). We claim that φ=ψγ\varphi=\psi_{\gamma}. We prove this by induction on the length of a walk representing a vertex in UU, starting with our base case length 0 walk (v)(v). Let α\alpha have a representative of length nn, of the form βe\beta*e for an edge ee and a length n1n-1 walk β\beta, and assume that φ(β)=ψγ(α)\varphi(\beta)=\psi_{\gamma}(\alpha). We know that ee defines an edge ded_{e} from β\beta to α\alpha in UU, and since φ\varphi is a graph homomorphism that commutes with ρ\rho we have an edge φ(de)=de\varphi(d_{e})=d^{\prime}_{e} from φ(β)\varphi(\beta) to φ(α)\varphi(\alpha). Therefore φ(α)=φ(β)de=γβde\varphi(\alpha)=\varphi(\beta)*d^{\prime}_{e}=\gamma*\beta*d^{\prime}_{e}. Uniqueness of path lifting ensures that γβde=γα\gamma*\beta*d^{\prime}_{e}=\gamma*\alpha. Thus φ=ψγ\varphi=\psi_{\gamma} and so Φ\Phi is surjective, completing the proof.

Given a subgroup SD(G)S\leq D(G), we define G~=U/S\widetilde{G}=U/S as the quotient graph where we identify vertices [α]=[sα][\alpha]=[s\alpha] and edges [sde]=[de][sd_{e}]=[d_{e}] for any sSs\in S. We can show that any cover is the quotient of UU by some subgroup of the deck transformation group D(G)D(G). Our proof uses groupoid coverings and Proposition 4.7. We start by showing that the resulting quotient is a homotopy cover.

Theorem 4.16.

Let GG be a graph with universal homotopy cover ρ:UG\rho:U\to G, and let SD(G)S\leq D(G). Define r:U/SGr:U/S\to G on vertices by r[α]=ρ(α)r[\alpha]=\rho(\alpha) and on edges r[de]=ρ(de)=er[d_{e}]=\rho(d_{e})=e. Then rr is a homotopy covering map.

Proof.

We first note that rr is well-defined on equivalence classes forming the vertices U/SU/S because for any sSD(G)s\in S\leq D(G), we have ρsα=ρα\rho s\alpha=\rho\alpha; and hence rr is surjective on vertices because ρ\rho is. The map rr is similarly well-defined on edges, and moreover since ρ(sde)=ρ(de)=e\rho(sd_{e})=\rho(d_{e})=e the edges [de][d_{e}] from [α][\alpha] to [β][\beta] are still in bijection with edges ee from their endpoints w,ww,w^{\prime} in GG. Thus rr induces a bijection from NE([α])N_{E}([\alpha]) to NE(w)N_{E}(w). Therefore rr is a cover as in Defintion 4.1.

We now show that rr induces a covering map on fundamental groupoids Π(r):Π(U/S)Π(G)\Pi(r):\Pi(U/S)\to\Pi(G), meaning that it is bijective on stars. By definition, rp=ρrp=\rho where pp is the projection map p:UU/Sp:U\to U/S. Functoriality ensures that Π(ρ)=Π(rp)=Π(r)Π(p)\Pi(\rho)=\Pi(rp)=\Pi(r)\Pi(p). Now Π(p)\Pi(p) is surjective on stars: given a walk ([d1][d2][dn])([d_{1}][d_{2}]\cdots[d_{n}]) in U/SU/S, we know that there are representatives did_{i} in UU and siSs_{i}\in S such that 1(di1)=0(sidi)\partial_{1}(d_{i-1})=\partial_{0}(s_{i}d_{i}). By applying the sis_{i} in turn to the end of the walk, we obtain a walk in UU that maps to ([d1][d2][dn])([d_{1}][d_{2}]\cdots[d_{n}]) by pp.

Since Π(p)\Pi(p) is surjective on stars and Π(ρ)=Π(r)Π(p)\Pi(\rho)=\Pi(r)\Pi(p) is bijective on stars, we conclude that Π(r)\Pi(r) is bijective on stars. By Proposition 4.7, rr is a homotopy covering map.

Proposition 4.17.

Let GG be a graph, and UU be its universal cover. Then any connected homotopy cover G~\widetilde{G} is of the form U/SG~U/S\cong\widetilde{G} for some subgroup SD(G)S\leq D(G).

Proof.

Suppose that f:G~Gf:\widetilde{G}\to G is a homotopy covering map. Theorem 4.10 states that UU is also the universal homotopy cover of G~\widetilde{G} and there exists a homotopy covering map ρ~:UG~\tilde{\rho}:U\to\widetilde{G} such that ρ=pf\rho=pf. Define S={sD(G):sρ~=ρ~}S=\{s\in D(G):s\tilde{\rho}=\tilde{\rho}\}. We wish to show that U/SG~U/S\cong\widetilde{G}. In fact, S=D(G~)S=D(\widetilde{G}), and thus it suffcies to consider the case S=D(G)S=D(G) and prove that U/D(G)GU/D(G)\cong G.

Let r:U/D(G)Gr:U/D(G)\to G be defined by r[u]=ρ(u)r[u]=\rho(u). Then rr is a homotopy covering map by Theorem 4.16, and hence surjective. We will show that it is also injective. Suppose that r[α]=w=r[β]r[\alpha]=w=r[\beta] for vertices [α],[β]U[\alpha],[\beta]\in U. Since α,β\alpha,\beta have the same endpoint, define γ=αβ1\gamma=\alpha*\beta^{-1}, giving a walk starting and ending at vv, and let s=ψγs=\psi_{\gamma} as in Definition 4.14. Then ψγD(G)\psi_{\gamma}\in D(G) and ψγ(β)=α\psi_{\gamma}(\beta)=\alpha and so [α]=[β][\alpha]=[\beta] in U/D(G)U/D(G). Therefore rr is a bijection of vertices, and also a cover which gives a bijection on neighbourhoods. So rr is also a bijection on edges and hence an isomorphism.

Observation 4.18.

Since all homotopy covers of GG are isomorphic to quotients of UU, Observation 4.9 ensures that any homotopy cover G~\widetilde{G} of a graph GG is isomorphic to a supergraph of a homotopy cover Gs~\widetilde{G_{s}} of GsG_{s}, where vertices [α],[β][\alpha],[\beta] in G~\widetilde{G} have the same number of parallel edges between them as r[α],r[β]r[\alpha],r[\beta] do.

We can use Propostion 3.8 to give a description of the universal homotopy cover of a reflexive graph as it relates to the universal homotopy cover of its unlooped subgraph.

Notation 4.19.

Let GG be a reflexive graph. Let f:CGuf:C\to G_{u} be a homotopy covering map of the unlooped graph GuG_{u}. We define CC_{\ell} to be the graph CC with a loop appended to a vertex α\alpha for each loop incident to the vertex f(α)f(\alpha) in the graph GG.

With this definition, we get the following.

Theorem 4.20.

Suppose that GG is a reflexive graph. We define the cover CC of its unlooped graph GuG_{u} by C=UvGu/TC=U_{v}G_{u}/T, the quotient of the universal cover of GuG_{u} by the the subgroup TD(Gu)Πvv(Gu)T\leq D(G_{u})\cong\Pi_{v}^{v}(G_{u}) generated by 3-cycles. Then the universal cover of GG satisfies

UvGC×K2.U_{v}G\cong C_{\ell}\times K_{2}.
Proof.

We know that UvGU_{v}G has vertices which are defined by walks from vv in Πv(G)\Pi_{v}(G). We apply the isomorphism of Proposition 3.8 to get an arrow in the groupoid Πv(Gu)/T×/2\Pi_{v}(G_{u})/T\times\operatorname{\mathbb{Z}}/2 starting at vv, giving an isomorphism between the vertices of UvGU_{v}G and C×K2C_{\ell}\times K_{2}. We then define the mapping UvGC×K2U_{v}G\to C_{\ell}\times K_{2} on edges. Any edge de(UvG)d_{e}\in(U_{v}G) corresponds to extending a walk α\alpha by a single edge eE(G)e\in E(G) to αe\alpha*e. In this case, the parity of α\alpha and αe\alpha*e are opposite, and so α\alpha and αe\alpha*e are mapped to (αu,i)(\alpha_{u},i) and ((αe)u,1i)((\alpha*e)_{u},1-i) in C×K2C_{\ell}\times K_{2}. If αu=(αe)u\alpha_{u}=(\alpha*e)_{u} then e=e=\ell was a loop, and we define the image of ded_{e} to be the edge ce=dc_{e}=d_{\ell} with 0(c)=(αu,i),1(c)=(αu,1i)\partial_{0}(c_{\ell})=(\alpha_{u},i),\partial_{1}(c_{\ell})=(\alpha_{u},1-i). Otherwise we have αue=(αe)u\alpha_{u}*e=(\alpha*e)_{u} and there is an edge cec_{e}, 0(ce)=(αu,i),1(ce)=(αue,1i)\partial_{0}(c_{e})=(\alpha_{u},i),\partial_{1}(c_{e})=(\alpha_{u}*e,1-i). We then map deced_{e}\mapsto c_{e}. It is straightforward to check that this gives a bijective map on edges in which de¯ce¯\overline{d_{e}}\mapsto\overline{c_{e}}. Thus we get an isomorphism of graphs.

Corollary 4.21.

Let GG be reflexive and C3C_{3}-free graph, and let UvGU_{v}G denote its universal homotopy cover. Then UvGUvGuK2U_{v}G\cong U_{v}G_{u}\square K_{2}.

Proof.

We know that UvGC×K2U_{v}G\cong C_{\ell}\times K_{2} and since GG is C3C_{3} free, UvGu=CU_{v}G_{u}=C and so the parity of a walk in CC is well-defined. Define a homomorphism Φ:C×K2CK2\Phi:C_{\ell}\times K_{2}\to C\square K_{2} by (γ,i)(γ,i+k)(\gamma,i)\to(\gamma,i+k) where kk is the length of γ\gamma (mod 2). Since γ\gamma and γe\gamma*e always have opposite parity, the vertices (γ,i)(\gamma,i) and (γe,1i)(\gamma*e,1-i) are always mapped to the same layer in the box product and this gives a graph homomorphism. It is straightforward to see that Φ\Phi is bijective on vertices and edges and thus is an isomorphism.

Example 4.22.

We consider the graph GG from Example 3.10. In this case, the cover C=UGu/TC=UG_{u}/T is a quotient of the universal homotopy cover by the triangle (axea)(axea). The result is drawn below, with covering map ff defined by vivv_{i}\to v:

b0b_{0}a0a_{0}c0c_{0}d0d_{0}e0e_{0}x0x_{0}b1b_{1}a1a_{1}c1c_{1}d1d_{1}e1e_{1}x1x_{1}e1e_{-1}x1x_{-1}a2a_{2}\cdots\cdotsUGu/TUG_{u}/T

Following Theorem 4.20, we obtain UG=C×K2UG=C_{\ell}\times K_{2} with vertices given by ordered pairs (α,i)(\alpha,i) where αV(C),i/2\alpha\in V(C),i\in\mathbb{Z}/2 and edges defined by (α,0)(β,1)(\alpha,0)\sim(\beta,1) for each edge from f(α)f(\alpha) to f(β)f(\beta) in GG.

(b1,0)(b_{1},0)(a1,0)(a_{1},0)(c1,0)(c_{1},0)(d1,0)(d_{1},0)(e1,0)(e_{1},0)(x1,0)(x_{1},0)(a1,1)(a_{1},1)(b1,1)(b_{1},1)(c1,1)(c_{1},1)(d1,1)(d_{1},1)(e1,1)(e_{1},1)(x1,1)(x_{1},1)(b0,0)(b_{0},0)(a0,0)(a_{0},0)(c0,0)(c_{0},0)(d0,0)(d_{0},0)(e0,0)(e_{0},0)(x0,0)(x_{0},0)(a0,1)(a_{0},1)(b0,1)(b_{0},1)(c0,1)(c_{0},1)(d0,1)(d_{0},1)(e0,1)(e_{0},1)(x0,1)(x_{0},1)(a2,0)(a_{2},0)(a2,1)(a_{2},1)\cdots\cdots\cdots\cdots\cdots\cdotsUGUG
Proposition 4.23.

Let GG be a reflexive graph. Then any homotopy cover of GG is related to a homotopy cover CC of the unlooped graph GuG_{u} that lifts 3-cycles to 3-cycles, and has one of the following forms:

  • the reflexive graph CC_{\ell}, or

  • an unlooped graph of the form C×K2C_{\ell}\times K_{2}.

Proof.

By Proposition 4.17, any homotopy cover G~\widetilde{G} of GG has the form G~UvG/S\widetilde{G}\cong U_{v}G/S for SD(G)S\leq D(G). And by Theorem 4.15 and Proposition 3.8, D(G)ΠvvGΠvvGu/T×/2D(G)\cong\Pi_{v}^{v}G\cong\Pi_{v}^{v}G_{u}/T\times\mathbb{Z}/2. So SS has the form S×/2S^{\prime}\times\operatorname{\mathbb{Z}}/2 or S×{0}S^{\prime}\times\{0\}, where SS^{\prime} is a subgroup of ΠvvGu\Pi_{v}^{v}G_{u} which contains TT. We can define the homotopy cover of the unlooped graph to be C=UGu/SC=UG_{u}/S^{\prime}. Then if SS is of the form S×/2S^{\prime}\times\operatorname{\mathbb{Z}}/2 we obtain a reflexive homotopy cover G~\widetilde{G} of the first form, and if SS is of the form S×{0}S^{\prime}\times\{0\} we obtain G~\widetilde{G} containing no loops of the second form.

Example 4.24.

Recall GG and its fundamental groupoid and universal cover from Examples 3.10 and 4.22. Below, we depict two 2-covers of GG, one reflexive and one not.

a0a_{0}b0b_{0}c0c_{0}d0d_{0}e0e_{0}x0x_{0}a1a_{1}b1b_{1}c1c_{1}d1d_{1}e1e_{1}x1x_{1}                        a0a_{0}b0b_{0}c0c_{0}d0d_{0}e0e_{0}x0x_{0}a1a_{1}b1b_{1}c1c_{1}d1d_{1}e1e_{1}x1x_{1}

To the left, we have U/(abcdexa)2×/2U/\langle(abcdexa)^{2}\rangle\times\operatorname{\mathbb{Z}}/2, and to the right U/(abcdexa)×{0}U/\langle(abcdexa)\rangle\times\{0\}.

We finish out our discussion of homotopy covers by showing that the they satisfy a general homotpy lifting property.

Theorem 4.25.

Let f:G~Gf:\widetilde{G}\to G be a homotopy covering map, and suppose we have ϕ,ψ:KG\phi,\psi:K\to G which are homotopic via a homotopy H:K×InGH:K\times I_{n}\to G. Suppose that we can lift ϕ\phi to a map ϕ~:KG~\tilde{\phi}:K\to\widetilde{G} such that ϕ=fϕ~\phi=f\circ\widetilde{\phi}. Then there is a lift H~:K×InG~\widetilde{H}:K\times I_{n}\to\widetilde{G} such that H~|K×{0}=ϕ~{\tilde{H}}|_{K\times\{0\}}=\widetilde{\phi} and fH~=Hf\circ\widetilde{H}=H. In particular, this will ensure that H~|K×{n}=ψ~{\tilde{H}}|_{K\times\{n\}}=\widetilde{\psi} is a lift of ψ\psi such that ϕ~ψ~\widetilde{\phi}\simeq\widetilde{\psi}.

G~fK×{0}ϕ~K×InH:ϕψH~G