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

Generalized Borsuk Graphs

Francisco Martinez-Figueroa
Abstract

Given a finite group GG acting freely on a compact metric space MM, and ε>0\varepsilon>0, we define the GG-Borsuk graph on MM by drawing edges xyx\sim y whenever there is a non-identity gGg\in G such that 𝕕(x,gy)ε\mathbbm{d}\left(x,gy\right)\leq\varepsilon. We show that when ε\varepsilon is small, its chromatic number is determined by the topology of MM via its GG-covering number, which is the minimum kk such that there is a closed cover M=F1FkM=F_{1}\cup\dots\cup F_{k} with Fig(Fi)=F_{i}\cap g(F_{i})=\emptyset for all gG{𝟙}g\in G\setminus\{\mathbbm{1}\}. We are interested in bounding this number. We give lower bounds using GG-actions on Hom-complexes, and upper bounds using a recursive formula on the dimension of MM. We conjecture that the true chromatic number coincides with the lower bound, and give computational evidence. We also study random GG-Borsuk graphs, which are random induced subgraphs. For these, we compute thresholds for ε\varepsilon that guarantee that the chromatic number is still that of the whole GG-Borsuk graph. Our results are tight (up to a constant) when the GG-index and dimension of MM coincide.

1 Introduction

Given ε>0\varepsilon>0 and d1d\geq 1, the Borsuk graph Bord(ε)\text{Bor}^{d}(\varepsilon) is the infinite graph with vertex set the unit dd-sphere 𝕊d\mathbb{S}^{d} and edges xyx\sim y whenever 𝕕𝕊d(x,y)πε\mathbbm{d}_{\mathbb{S}^{d}}\left(x,y\right)\geq\pi-\varepsilon, that is, if the two points are ε\varepsilon-almost antipodal. It is well known that when ε\varepsilon is sufficiently small, its chromatic number is d+2d+2, which follows from the topology of 𝕊d\mathbb{S}^{d} via Borsuk-Ulam’s Theorem, and an explicit (d+2)(d+2) coloring given by projecting the facets of an inscribed regular (d+1)(d+1)-simplex (see e.g. [Mat08, KMF20]).

The Borsuk graph has been studied in relation to Borsuk’s conjecture and distance graphs (see e.g. [Rai12, PRS17, Sag18, BM14]). It is also mentioned in Lovász’s proof of Kneser’s conjecture [Lov78] where he initiated the study of topological obstructions to the chromatic number of graphs. This study has been continued, among others, by Babson’s and Kozlov’s work on the topology of Hom-complexes of graphs [BK03, BK06, Koz08]. In [Kah07], Kahle computed that these topological lower bounds are not efficient for the chromatic number of Erdős–Rényi random graphs. In contrast, in [KMF20] we showed that the topology of the underlying dd-sphere still dictates the chromatic number of random Borsuk graphs, and we described the threshold of ε\varepsilon for this to happen.

Regarding the antipodal map on a sphere as an action of the cyclic group of order two 2={𝟙,ν}\mathbb{Z}_{2}=\{\mathbbm{1},\nu\}, we can equivalently define the Borsuk graph Bord(ε)\text{Bor}^{d}(\varepsilon) as the graph with vertices 𝕊d\mathbb{S}^{d}, and edges xyx\sim y whenever 𝕕𝕊d(x,ν(y))ε\mathbbm{d}_{\mathbb{S}^{d}}\left(x,\nu(y)\right)\leq\varepsilon. This formulation naturally generalizes the Borsuk graphs to any metric space where a group GG acts. Thus, if GG is a group acting on a metric space MM, and ε>0\varepsilon>0 is given, the GG-Borsuk graph on MM, denoted 𝒢G(M,ε)\mathcal{G}_{G}\left(M,\varepsilon\right), is the graph with vertex set MM, and edges xyx\sim y whenever there exist gG,g𝟙g\in G,g\neq\mathbbm{1} s.t. 𝕕M(x,gy)ε\mathbbm{d}_{M}\left(x,gy\right)\leq\varepsilon. In this paper, we study the chromatic number of such GG-Borsuk graphs, when GG is a finite group and the underlying space MM is finitely triangulable.

Just as with regular Borsuk graphs, the chromatic number of GG-Borsuk graphs is determined by the topology of the underlying space, however accurately bounding these requires more work. To get lower bounds, we study the corresponding hom-complexes as GG-spaces, applying standard tools for GG-spaces as described in [Mat08, Ch. 6]. We also establish some upper bounds, however this are far away from the lower bounds. Computer-aided examples suggest that the topological lower bounds might actually tight.

We also generalize the random Borsuk graphs studied in [KMF20]: a random GG-Borsuk graph on MM, is the induced subgraph by taking nn uniform i.i.d. random points on MM. For these, we exhibit threshold for ε\varepsilon such that asymptotically almost surely the chromatic number of the random subgraph coincides with that of the whole GG-Borsuk graph.

This paper is organized as follows. In section 2, we summarize the necessary background on GG-spaces and GG-simplicial complexes needed for the following sections. We also recall the definition of graph homomorphisms, and describe some conventions that we use throughout this paper. In section 3, we summarize our definitions and theorems about GG-Borsuk graphs. In section 4, we prove the topological lower bound for the chromatic number. In section 5, we prove our upper bound. In section 6, we study the random GG-Borsuk graphs. Finally, we include an appendix where we describe our approach to produce computer-aided examples.

2 Background

In this section we recall the standard definitions and properties of GG-spaces and GG-simplicial complexes, including EdGE_{d}G spaces and the GG-index. Our exposition follows very closely that of Matoušek in [Mat08, Ch. 6]. We also include the definition of graph homomorphisms and their relation to graph colorings (see more at [HN04, Ch. 1]). In addition, we make some useful conventions for the remainder of this paper.

Definition 2.1 (GG-spaces).

Given a finite group GG, a topological space MM, is a 𝐆\bm{G}-space if GG acts on MM via continuous maps. We say it is a free 𝐆\bm{G}-space if the GG-action is free, that is, if for all gGg\in G, g𝟙g\neq\mathbbm{1} the map g:MMg:M\to M has no fixed points.

Definition 2.2 (Simplicial GG-spaces).

Let GG be a finite group.

  • A simplicial complex KK is a 𝑮\bm{G}-simplicial complex if GG acts on KK via simplicial maps. We say the action is free, if for every face σK\sigma\in K and every gGg\in G with g𝟙g\neq\mathbbm{1}, we have σgσ=\sigma\cap g\sigma=\emptyset.

  • Moreover, a geometric realization of KK, is a geometric 𝑮\bm{G}-simplicial complex, if GG acts on it via linear extensions of simplicial maps. In this case, we consider K\|K\| as a metric space with the shortest path distance. Note that the GG-action is indeed continuous with this metric. (Here, and throughout this paper, \|\cdot\| denotes a fixed geometric realization).

In this paper: we will only consider free G\bm{G}-spaces and free G\bm{G}-simplicial complexes, even if we don’t say this explicitly.

Note that KK being a free GG-simplicial complex is a stronger condition than just asking it to be free on the geometric realization. However, if KK is a compact triangulable space and GG acts on it freely, we can always find a small enough triangulation such that the simplicial approximation theorem produces a free GG-simplicial complex.

We will mostly focus on compact metric GG-spaces where the actions are via Lipschitz maps or isometries. Note that when KK is a geometric GG-simplicial complex, GG acts on it via piece-wise linear maps. Thus, if KK has a finite number of faces, the GG-action is automatically Lipschitz.

Definition 2.3 (GG-maps).

Let GG be a finite group.

  1. 1.

    Let XX and YY be GG-spaces. A map φ:XY\varphi:X\to Y is a 𝑮\bm{G}-map if φ(g(x))=g(φ(x))\varphi(g(x))=g(\varphi(x)) for all xXx\in X, and all gGg\in G.

  2. 2.

    Let KK and MM be GG-simplicial complexes. A simplicial map φ:KM\varphi:K\to M is a GG-map if φ(gv)=gφ(v)\varphi(gv)=g\varphi(v) for all vV(K)v\in V(K) and all gGg\in G.

Notation: In both cases we denote that φ\varphi is a GG-map by writing φ:X𝐺Y\varphi:X\xrightarrow{G}Y. We will also write just X𝐺YX\xrightarrow{G}Y, to mean that there exists some GG-map from XX to YY. It is straightforward to check that composition of GG-maps produces a GG-map and that the identity is trivially a GG-map. More generally, GG-spaces with GG-maps form a category, and (geometric) GG-simplicial complexes with GG-maps form a category as well.

When G=2G=\mathbb{Z}_{2}, the canonical example of a 2\mathbb{Z}_{2}-space is the dd-sphere 𝕊d\mathbb{S}^{d} with the antipodal map. For examples of 2\mathbb{Z}_{2}-simplicial complexes, we can take any triangulation of 𝕊d\mathbb{S}^{d} that is symmetrical with respect to the origin. In particular, the cross-polytopes d\diamond^{d} provide a family of such examples. Having one canonical 2\mathbb{Z}_{2}-space for each dimension, makes the spheres the perfect benchmark for comparing with other 2\mathbb{Z}_{2}-spaces. The idea of benchmark spaces can be extended to other groups via the finite classifying spaces EdGE_{d}G, here again we follow the notation of [Mat08].

Definition 2.4 (Finite Classifying Spaces).

Let GG be a finite group, and d0d\geq 0. A finite classifying space of dimension 𝐝\bm{d}, denoted EdGE_{d}G, is

  1. 1.

    a finite free GG-simplicial complex,

  2. 2.

    dd-dimensional, and

  3. 3.

    (d1)(d-1)-connected

This definition doesn’t rule out the existence of different EdGE_{d}G classifying spaces for the same group GG and dimension dd. However, they are always GG-equivalent in the sense that if both XX and YY are EdGE_{d}G spaces, then we always have X𝐺Y\|X\|\xrightarrow{G}Y and Y𝐺X\|Y\|\xrightarrow{G}X [Mat08, Lemma 6.2.2 ]. Let us give some useful examples of EdGE_{d}G spaces.

Examples 2.5.
  1. 1.

    For G=2G=\mathbb{Z}_{2}, any symmetrical triangulation of 𝕊d\mathbb{S}^{d} is a Ed2E_{d}\mathbb{Z}_{2} space. In particular, the dd-dimensional cross-polytope d\diamond^{d}, which can be described via simplicial joins as

    d=(𝕊0)(d+1)=𝕊0𝕊0d+1 times.\diamond^{d}=(\mathbb{S}^{0})^{*(d+1)}=\underbrace{\mathbb{S}^{0}*\cdots*\mathbb{S}^{0}}_{d+1\text{ times}}.
  2. 2.

    By some abuse of notation, we can regard a finite group GG as a 0-dimensional simplicial complex with its vertices labeled by the elements of the group. Then, for any d0d\geq 0, GG acts freely on the space G(d+1)G^{*(d+1)} via left-multiplication on each coordinate of the simplicial join. By elementary properties of the simplicial join operation, G(d+1)G^{*(d+1)} is (d1)(d-1)-connected and dd-dimensional, so it is an EdGE_{d}G classifying space. Also, by the symmetries of this construction, it can be seen that it can be realized by a geometric simplicial complex where GG acts via isometries.

  3. 3.

    Every cyclic group m\mathbb{Z}_{m} acts freely on 𝕊1\mathbb{S}^{1} via (2π)/m(2\pi)/m rotations. Since 𝕊1\mathbb{S}^{1} is 1-dimensional and connected, it gives a E1mE_{1}\mathbb{Z}_{m} space.

  4. 4.

    Following the previous item, each cyclic group m\mathbb{Z}_{m} acts freely on the space m𝕊1\mathbb{Z}_{m}*\mathbb{S}^{1}, which is simply-connected and 2-dimensional, so it is a EdmE_{d}\mathbb{Z}_{m} space. We can understand this space as a collection of mm cones over the same 𝕊1\mathbb{S}^{1}. Pictorially, we represent these spaces by showing each v𝕊1v*\mathbb{S}^{1} as an independent disk, where the boundaries must be identified. We use this depiction in Figure 1(d), where we draw the disks at the vertices of a regular mm-gon, so that the m\mathbb{Z}_{m} action can be observed by rotating the picture by multiples of 2π/m2\pi/m.

Comparing any GG-space with the finite classifying spaces, the GG-index is defined as follows.

Definition 2.6 (GG-index).

For a GG-space MM, its 𝐆\bm{G}-index is defined as

indG(M):=min{d:M𝐺EdG}\text{ind}_{G}\left(M\right):=\min\{d:M\xrightarrow[]{G}E_{d}G\}

We summarize some properties of the GG-index.

Theorem 2.7 ([Mat08, Prop. 6.2.4]).
  1. 1.

    If X𝐺YX\xrightarrow{G}Y then indG(X)indG(Y)\text{ind}_{G}\left(X\right)\leq\text{ind}_{G}\left(Y\right).

  2. 2.

    indG(EdG)=d\text{ind}_{G}\left(E_{d}G\right)=d for all d0d\geq 0.

  3. 3.

    indG(XY)indG(X)+indG(Y)+1\text{ind}_{G}\left(X*Y\right)\leq\text{ind}_{G}\left(X\right)+\text{ind}_{G}\left(Y\right)+1.

  4. 4.

    If XX is kk-connected, then indG(X)k+1\text{ind}_{G}\left(X\right)\geq k+1.

  5. 5.

    If KK is a free simplicial GG-complex of dimension dd, then indG(X)d\text{ind}_{G}\left(X\right)\leq d.

Note item 2 is a generalization of Borsuk–Ulam’s Theorem, and guarantees the GG-index is well defined.

Throughout this paper whenever we say HH is a graph, we mean it is a finite, simple, loop-less graph. That is, its vertex set V(H)V(H) is a finite set, and its edges E(H)E(H) are unordered pairs of vertices. We will also use the notation uvu\sim v between vertices of HH, to represent the edge {u,v}E(H)\{u,v\}\in E(H). We will denote the complete graph on mm vertices by KmK_{m}. The next couple of definitions are meant as a quick review of graph homomorphisms and colorings, for a more detailed exposition see [HN04, Ch. 1].

Definition 2.8.

Given two graphs H1H_{1} and H2H_{2}, a graph homomorphism from H1H_{1} to H2H_{2} is a function f:V(H1)V(H2)f:V(H_{1})\to V(H_{2}) defined on their vertices such that whenever uvu\sim v in H1H_{1} then also f(u)f(v)f(u)\sim f(v) in H2H_{2}.

We can thus understand graph homomorphisms as functions that send vertices to vertices and edges to edges. We will denote them as f:H1H2f:H_{1}\to H_{2}. Again, we may just write H1H2H_{1}\to H_{2} to mean that there exists some graph homomorphism from H1H_{1} to H2H_{2}. It is known that the collection of graphs with graph homomorphisms form a category (see [Koz08, HN04]).

Given two graphs H1H_{1} and H2H_{2}, we will denote the set of graph homomorphisms from H1H_{1} to H2H_{2} as Hom(H1,H2)\text{Hom}(H_{1},H_{2}). Babson and Kozlov [BK03, BK06] introduced a canonical way of endowing Hom(H1,H2)\text{Hom}(H_{1},H_{2}) with a topological structure. We will discuss more about it in section 5.

Graph colorings and chromatic number can also be described via graph homomorphisms.

Definition 2.9.

Given a graph HH, an m-coloring is a graph homomorphism φ:HKm\varphi:H\to K_{m}. The chromatic number of HH, is the least mm such that an mm-coloring exits, i.e.

χ(H):=min{m:Hom(H,Km)}.\chi(H):=\min\{m:\text{Hom}(H,K_{m})\neq\emptyset\}.

3 Definitions and Results

We start by giving a definition that generalizes Borsuk graphs on spheres to any metric GG-space.

Definition 3.1 (GG-Borsuk Graphs).

Let GG be a finite group. For a metric GG-space MM, XMX\subset M, and ε>0\varepsilon>0 we define the 𝐆\bm{G}-Borsuk Graph 𝒢G(X,ε)\mathcal{G}_{G}\left(X,\varepsilon\right) as the graph with vertices V(𝒢G(X,ε))=XV(\mathcal{G}_{G}\left(X,\varepsilon\right))=X and edges xyx\sim y whenever there is a non-identity element gGg\in G such that 𝕕M(x,gy)ε\mathbbm{d}_{M}\left(x,gy\right)\leq\varepsilon.

Thus the 2\mathbb{Z}_{2}-Borsuk Graph 𝒢2(𝕊d,ε)\mathcal{G}_{\mathbb{Z}_{2}}\left(\mathbb{S}^{d},\varepsilon\right), where 2\mathbb{Z}_{2} acts on 𝕊d\mathbb{S}^{d} via the antipodal map, is the Borsuk Graph on 𝕊d\mathbb{S}^{d}. It is well known that the chromatic number of the 2\mathbb{Z}_{2}-Borsuk Graph on the dd-Sphere is d+2d+2, which follows from Borsuk–Ulam’s Theorem (see e.g. [Mat08]). More specifically, it depends on Lyusternik–Schnirerlman’s Theorem, which is equivalent to Borsuk–Ulam’s, concerning the minimum number of open sets covering 𝕊d\mathbb{S}^{d} s.t. they don’t contain pairs of antipodal points. In this spirit, we define the GG-covering number of a GG-metric space, and study its relation with GG-Borsuk graphs.

Definition 3.2 (GG-covering number).

Let GG be a finite group. Let MM be a free GG-space. Then a (open) 𝐆\bm{G}-cover of MM is an open cover M=U1UkM=U_{1}\cup\dots\cup U_{k}, such that UigUi=U_{i}\cap gU_{i}=\emptyset for all i=1,,ki=1,\dots,k and all gG,g𝟙g\in G,g\neq\mathbbm{1}. Then the (open) 𝐆\bm{G}-covering number of MM is:

covG(M)=min{k: there exists a G-cover M=U1Uk}\text{cov}_{G}\left(M\right)=\min\{k:\text{ there exists a }G\text{-cover }M=U_{1}\cup\dots\cup U_{k}\}

Similarly, we define closed GG-covers and the closed 𝐆\bm{G}-covering number of MM, covG(M)¯\overline{\text{cov}_{G}\left(M\right)}, using closed covers instead.

Notation:

Theorem 4.1 will show that the open and closed covering numbers coincide for compact spaces, and that they are invariant under GG-equivalence. Since all EdGE_{d}G spaces are GG-equivalent, the GG-covering number covG(EdG)\text{cov}_{G}\left(E_{d}G\right) is independent of the specific classifying space, so we will denote it simply as covG(d)\text{cov}_{G}\left(d\right).

The GG-covering number the chromatic number of GG-Borsuk graphs on the same space are related via the next result.

Theorem 3.3.

If MM is a compact metric space and GG acts via Lipschitz maps. Then there exist constants DD and ε0\varepsilon_{0} such that for any 0<ε<ε00<\varepsilon<\varepsilon_{0} and any XMX\subset M a (ε/D)(\varepsilon/D)-net of MM, we have

χ(𝒢G(X,ε))=covG(M).\chi\left(\mathcal{G}_{G}\left(X,\varepsilon\right)\right)=\text{cov}_{G}\left(M\right).

This relationship is what motivates us to find bounds for the GG-covering number, and, at the same time this duality is one of our main tools finding these bounds.

Theorem 3.4.

Let GG be a finite group and KK a geometric GG-simplicial complex. Let k=indG(K)k=\text{ind}_{G}\left(K\right). Then

k+|G|covG(K)k(|G|1)+2.k+|G|\leq\text{cov}_{G}\left(K\right)\leq k(|G|-1)+2.

Thus we get linear upper and lower bounds for covG(K)\text{cov}_{G}\left(K\right), depending only on the size of GG and the GG-index of KK. In the case G=2G=\mathbb{Z}_{2} we get the tight result covG(K)=ind2(K)+2\text{cov}_{G}\left(K\right)=\text{ind}_{\mathbb{Z}_{2}}\left(K\right)+2, where the lower bound is a generalization of Lyusternik–Schnirelmann–Borsuk Theorem, and the upper bound is obtained by “coloring” the facets of a regular (k+1)(k+1)-dimensional simplex inscribed in 𝕊k\mathbb{S}^{k} (see e.g.[KMF20, Mat02]).

In general, we conjecture that the lower bound gives the true chromatic number. If KK has GG-index kk, then K𝐺EkGK\xrightarrow{G}E_{k}G, and so by Theorem 4.1, covG(K)covG(EkG)=covG(k)\text{cov}_{G}\left(K\right)\leq\text{cov}_{G}\left(E_{k}G\right)=\text{cov}_{G}\left(k\right), so we only need to upper bound the GG-covering number for EdGE_{d}G spaces. Thus, we state our conjecture as follows.

Conjecture 3.5.

Let GG be a finite group. Then covG(d)=|G|+d\text{cov}_{G}\left(d\right)=|G|+d.

We also state the following less general conjectures, which may hold even if Conjecture 3.5 results to be false.

Conjecture 3.6.
  1. 1.

    For m2m\geq 2 and d0d\geq 0,

    covm(d)=m+d.\text{cov}_{\mathbb{Z}_{m}}\left(d\right)=m+d.
  2. 2.

    Let GG be a finite group. Then covG(d)<covG(d+1)\text{cov}_{G}\left(d\right)<\text{cov}_{G}\left(d+1\right) for all d0d\geq 0.

Notice that item 2 in the conjecture asks if covG()\text{cov}_{G}\left(\cdot\right) is strictly increasing, since the fact that it is weakly monotone follows trivially from the GG-map EdG=G(d+1)𝐺G(d+2)=Ed+1GE_{d}G=G^{*(d+1)}\xhookrightarrow{G}G^{*(d+2)}=E_{d+1}G, and Theorem 4.1.

Table 1 summarizes the cases for which we have been able to compute covG(d)\text{cov}_{G}\left(d\right) exactly. Notice that for all of these, Conjecture 3.5 holds. Specific m\mathbb{Z}_{m}-coverings for E2mE_{2}\mathbb{Z}_{m} when m=3,4,5m=3,4,5 and 6 are shown in Figure 1(d). Here we use m𝕊1\mathbb{Z}_{m}*\mathbb{S}^{1} as E2mE_{2}\mathbb{Z}_{m} space, and is depicted as described in Examples 2.5. These GG-covers were found solving integer linear programming problems with the free software [SageMath] and a student license of [Gurobi]. We further explain our approach in the appendix.

𝑮\bm{G}-Index Group Covering Number
𝒅\bm{d} 𝑮\bm{G} cov(d)G\bm{{}_{G}(d)}
any d0d\geq 0 2\mathbb{Z}_{2} d+2d+2
0 any GG |G||G|
1 any GG |G|+1|G|+1
2 3\mathbb{Z}_{3} 5
4\mathbb{Z}_{4} 6
2×2\mathbb{Z}_{2}\times\mathbb{Z}_{2} 6
5\mathbb{Z}_{5} 7
6\mathbb{Z}_{6} 8
3 3\mathbb{Z}_{3} 6
Table 1: Known Covering Numbers
Refer to caption
(a) A cover of E23E_{2}\mathbb{Z}_{3} with 5 colors.
Refer to caption
(b) A cover of E24E_{2}\mathbb{Z}_{4} with 6 colors.
Refer to caption
(c) A cover of E25E_{2}\mathbb{Z}_{5} with 7 colors.
Refer to caption
(d) A cover of E26E_{2}\mathbb{Z}_{6} with 8 colors.
Figure 1: Optimal GG-coverings for 2-dimensional classifying spaces of G=3,4,5G=\mathbb{Z}_{3},\mathbb{Z}_{4},\mathbb{Z}_{5} and 6\mathbb{Z}_{6}

.

The lower bound in Theorem 3.4, provides a generalization of the Lyusternik–Schnirelmann–Borsuk Theorem to GG-spaces.

Corollary 3.7 (Generalization of LSB-Theorem).

Let GG be a finite group and KK a finite GG-simplicial complex. Suppose MM is dd-connected and |G|=m|G|=m. If M=U1Ud+mM=U_{1}\cup\dots\cup U_{d+m} is an open (or closed) cover of KK, then there exist xKx\in K, an index 1id+m1\leq i\leq d+m and an element gG{𝟙}g\in G\setminus\{\mathbbm{1}\} such that x,g(x)Uix,g(x)\in U_{i}.

Proof.

By Theorem 2.7, indG(K)d+1\text{ind}_{G}\left(K\right)\geq d+1, and by Theorem 3.4, covG(K)indG(K)+|G|m+d+1\text{cov}_{G}\left(K\right)\geq\text{ind}_{G}\left(K\right)+|G|\geq m+d+1. Thus there are not open (nor closed, by Theorem 4.1) GG-covers of KK. ∎

We also generalize the random Borsuk graphs studied in [KMF20] to random induced subgraphs of GG-Borsuk graphs.

Definition 3.8 (Random GG-Borsuk graphs).

Let GG be a finite group, and MM a compact metric GG-space. We define the random GG-Graph on MM, denoted by 𝒢G(n,ε)\mathcal{G}_{G}\left(n,\varepsilon\right), as the GG-Borsuk graph 𝒢G(X(n),ε)\mathcal{G}_{G}\left(X(n),\varepsilon\right), where X(n)={x1,,xn}X(n)=\{x_{1},\dots,x_{n}\} is a set of nn i.i.d. uniform random points on MM.

In the same spirit as Theorems 1.2 and 1.3 of [KMF20], the next result provides a tight threshold (up to a constant) for ε\varepsilon, such that the random GG-Borsuk graphs have the chromatic number equal to the GG-covering number of its dimension. In this theorem, and throughout this paper, we will say that a random event AnA_{n} happens asymptotically almost surely (a.a.s.) if [An]1\mathbb{P}\left[A_{n}\right]\to 1 as nn\to\infty.

Theorem 3.9.

Let GG be a finite group and KK a geometric EdGE_{d}G space. Then, there exist constants C1C_{1} and C2C_{2} (independent from nn), such that:

  1. 1.

    If εC1(lognn)1/d\varepsilon\geq C_{1}\left(\frac{\log n}{n}\right)^{1/d}, then a.a.s. χ(𝒢G(n,ε))=covG(d).\chi(\mathcal{G}_{G}\left(n,\varepsilon\right))=\text{cov}_{G}\left(d\right).

  2. 2.

    If εC2(lognn)1/d\varepsilon\leq C_{2}\left(\frac{\log n}{n}\right)^{1/d}, then a.a.s. χ(𝒢G(n,ε))covG(d1).\chi(\mathcal{G}_{G}\left(n,\varepsilon\right))\leq\text{cov}_{G}\left(d-1\right).

Instead of proving Theorem 3.9, we will prove the more general Theorem 7.2, which can be applied to any geometric GG-simplicial complex, even when the GG-index and dimension don’t coincide. However, in such general cases, the thresholds found don’t coincide.

A nice application of GG-Borsuk graphs, is that they can produce graphs with large chromatic number and prescribed clique number. A similar construction using GG-actions can also be found in [Dan18].

Corollary 3.10.

Given mtm\leq t integers, there is a graph HH such that ω(H)=m\omega(H)=m and χ(H)t\chi(H)\geq t.

Proof.

Let GG be any group of order mm, and let d=tmd=t-m. Consider KK a geometric EdGE_{d}G space. Let ε1\varepsilon_{1} and D1D_{1} be the corresponding constants given by Theorem 4.2, and ε2\varepsilon_{2}, D2D_{2}, the ones given by Lemma 4.3. Let ε0=min{ε1,ε2}\varepsilon_{0}=\min\{\varepsilon_{1},\varepsilon_{2}\} and D=max{D1,D2}D=\max\{D_{1},D_{2}\}. Take any XEdGX\subset E_{d}G, an (ε/D)(\varepsilon/D)-net, for some ε<ε0\varepsilon<\varepsilon_{0}. Take H=𝒢G(X,ε)H=\mathcal{G}_{G}\left(X,\varepsilon\right). Then, by Lemma 4.3, ω(H)=m\omega(H)=m. And by Theorems 4.2 and 3.4,

χ(H)=covG(d)|G|+d=t.\chi(H)=\text{cov}_{G}\left(d\right)\geq|G|+d=t.

4 Properties of cov𝑮\bm{\text{cov}_{G}}

In this section we state and prove the main properties of covG()\text{cov}_{G}\left(\cdot\right) and its relationship with the chromatic number of GG-Borsuk graphs. The following Theorem summarizes a few of these properties.

Theorem 4.1.

Let GG be a finite group, and MM and KK be GG-spaces. Then:

  1. 1.

    If M𝐺KM\xrightarrow[]{G}K, then covG(M)covG(K)\text{cov}_{G}\left(M\right)\leq\text{cov}_{G}\left(K\right). In particular, if MM and KK are GG-equivalent, covG(M)=covG(K)\text{cov}_{G}\left(M\right)=\text{cov}_{G}\left(K\right).

  2. 2.

    If φ:M𝐺K\varphi:M\xrightarrow{G}K is a Lipschitz map with constant λ\lambda, then for any XMX\subset M, and any ε>0\varepsilon>0, it induces a graph homomorphism

    φ:𝒢G(X,ε)𝒢G(φ(X),λε).\varphi:\mathcal{G}_{G}\left(X,\varepsilon\right)\to\mathcal{G}_{G}\left(\varphi(X),\lambda\varepsilon\right).
  3. 3.

    If MM is a compact metric space and GG-acts via Lipschitz maps, then covG(M)=covG(M)¯.\text{cov}_{G}\left(M\right)=\overline{\text{cov}_{G}\left(M\right)}. So the covering number can be computed using either open or closed covers.

  4. 4.

    If GGG^{\prime}\leq G is a subgroup, then covG(M)covG(M)\text{cov}_{G^{\prime}}\left(M\right)\leq\text{cov}_{G}\left(M\right)

Proof.
  1. 1.

    Let k=covG(K)k=\text{cov}_{G}\left(K\right), we may assume k<k<\infty otherwise there’s nothing to prove. Thus there is an open GG-cover K=U1UkK=U_{1}\cup\dots\cup U_{k} such that Uig(Uj)=U_{i}\cap g(U_{j})=\emptyset for all ii and all gG{𝟙}g\in G\setminus\{\mathbbm{1}\}. Let f:M𝐺Kf:M\xrightarrow[]{G}K, then M=f1(U1)f1(Uk)M=f^{-1}(U_{1})\cup\dots\cup f^{-1}(U_{k}) gives an open cover for MM. Moreover, for each ii, and gG{𝟙}g\in G\setminus\{\mathbbm{1}\}, we have

    f((f1(Ui)g(f1(Ui)))Uig(Ui)=f(\left(f^{-1}(U_{i})\cap g(f^{-1}(U_{i}))\right)\subset U_{i}\cap g(U_{i})=\emptyset

    so the sets f1(Ui)f^{-1}(U_{i}) are an open GG-cover for MM. Therefore, covG(M)k=covG(K)\text{cov}_{G}\left(M\right)\leq k=\text{cov}_{G}\left(K\right).

  2. 2.

    We just need to check that φ\varphi maps edges to edges. Take xyx\sim y in 𝒢G(X,ε)\mathcal{G}_{G}\left(X,\varepsilon\right), so x,yXx,y\in X and there is some gG,g𝟙g\in G,g\neq\mathbbm{1} s.t. 𝕕M(x,gy)ε\mathbbm{d}_{M}\left(x,gy\right)\leq\varepsilon. Since φ\varphi is Lipschitz with constant λ\lambda, then

    𝕕K(φ(x),gφ(y))=𝕕K(x,φ(gy))λ𝕕M(x,gy)λε,\mathbbm{d}_{K}\left(\varphi(x),g\varphi(y)\right)=\mathbbm{d}_{K}\left(x,\varphi(gy)\right)\leq\lambda\mathbbm{d}_{M}\left(x,gy\right)\leq\lambda\varepsilon,

    and so φ(x)φ(y)\varphi(x)\sim\varphi(y) in 𝒢G(φ(X),λε)\mathcal{G}_{G}\left(\varphi(X),\lambda\varepsilon\right). Thus φ\varphi induces a graph homomorphism.

  3. 3.

    We may assume both covG(M)\text{cov}_{G}\left(M\right) and covG(M)¯\overline{\text{cov}_{G}\left(M\right)} are finite. If one of them is finite, the following proof gives that the other must be finite as well. Otherwise, they are both infinite.

    • Let k=covG(M)¯k=\overline{\text{cov}_{G}\left(M\right)}. Write M=F1FkM=F_{1}\cup\dots\cup F_{k}, where the FiF_{i}’s are a closed GG-cover. Note Fi×FiF_{i}\times F_{i} is compact, and for each g𝟙g\neq\mathbbm{1}, the map 𝕕(,g()):Fi×Fi+\mathbbm{d}\left(\cdot,g(\cdot)\right):F_{i}\times F_{i}\to\mathbb{R}^{+} is continuous (it is indeed positive, because Fig(Fi)=F_{i}\cap g(F_{i})=\emptyset). Since bi,g:=min{𝕕(x,g(y)):x,yFi}b_{i,g}:=\min\{\mathbbm{d}\left(x,g(y)\right):x,y\in F_{i}\} is attained, it must be that bi,g>0b_{i,g}>0. Let b=min{bi,g:1ik,gG,g𝟙}b=\min\{b_{i,g}:1\leq i\leq k,g\in G,g\neq\mathbbm{1}\}, note b>0b>0.

      Since GG acts via Lipschitz maps, let λg\lambda_{g} be the corresponding Lipschitz constant for the map gG{𝟙}g\in G\setminus\{\mathbbm{1}\}. Let D=1+max{λg:gG,g𝟙}D=1+\max\{\lambda_{g}:g\in G,g\neq\mathbbm{1}\}. For each i=1,,ki=1,\dots,k define Ui=xFiint(B(x,b/D))U_{i}=\bigcup_{x\in F_{i}}\text{int}\left(B(x,b/D)\right). Thus the UiU_{i}’s are open, and since FiUiF_{i}\subset U_{i}, they also cover MM. Now we prove the UiU_{i}’s are a GG-cover.

      Suppose by way of contradiction that yUig(Ui)y\in U_{i}\cap g(U_{i}). So there is yUiy^{\prime}\in U_{i} with y=g(y)y=g(y^{\prime}). By definition of UiU_{i}, there are x,xFix,x^{\prime}\in F_{i} such that 𝕕(x,y)<b/D\mathbbm{d}\left(x,y\right)<b/D and 𝕕(x,y)<b/D\mathbbm{d}\left(x^{\prime},y^{\prime}\right)<b/D. Then

      bbi,g𝕕(x,g(x))\displaystyle b\leq b_{i,g}\leq\mathbbm{d}\left(x,g(x^{\prime})\right) 𝕕(x,y)+𝕕(y,g(x))=𝕕(x,y)+𝕕(g(y),g(x))\displaystyle\leq\mathbbm{d}\left(x,y\right)+\mathbbm{d}\left(y,g(x^{\prime})\right)=\mathbbm{d}\left(x,y\right)+\mathbbm{d}\left(g(y^{\prime}),g(x^{\prime})\right)
      𝕕(x,y)+λg𝕕(y,x)<bD+λgbD=1+λgDbb\displaystyle\leq\mathbbm{d}\left(x,y\right)+\lambda_{g}\mathbbm{d}\left(y^{\prime},x^{\prime}\right)<\frac{b}{D}+\lambda_{g}\frac{b}{D}=\frac{1+\lambda_{g}}{D}b\leq b

      But b<bb<b is a contradiction, so Uig(Ui)=U_{i}\cap g(U_{i})=\emptyset for all ii, g𝟙g\neq\mathbbm{1}. Therefore, covG(M)covG(M)¯\text{cov}_{G}\left(M\right)\leq\overline{\text{cov}_{G}\left(M\right)}.

    • Let k=covG(M)k=\text{cov}_{G}\left(M\right). Write M=U1UkM=U_{1}\cup\dots\cup U_{k}, where the UiU_{i}’s are an open GG-cover. For each xMx\in M, choose a superset xUix\in U_{i}, and let FxF_{x} be a closed neighborhood of xx contained in UiU_{i}. The collection {interior(Fx):xM}\{\text{interior}(F_{x}):x\in M\} is an open cover of MM, by compactness, there is a finite subcover 𝒞\mathcal{C}.

      For each i=1,,ki=1,\dots,k, let Fi=FxF_{i}=\bigcup F_{x} taking the union over the closed sets such that interior(Fx)𝒞\text{interior}(F_{x})\in\mathcal{C} and FxUiF_{x}\subset U_{i}. Thus F1,,FkF_{1},\dots,F_{k} is a closed cover of MM. Since FiUiF_{i}\subset U_{i}, trivially Fig(Fi)=F_{i}\cap g(F_{i})=\emptyset for all ii, gGg\in G, g𝟙g\neq\mathbbm{1}. Therefore covG(M)¯covG(M)\overline{\text{cov}_{G}\left(M\right)}\leq\text{cov}_{G}\left(M\right).

  4. 4.

    Note any GG-cover of MM, will also be a GG^{\prime}-cover of MM, so trivially covG(M)covG(M)\text{cov}_{G^{\prime}}\left(M\right)\leq\text{cov}_{G}\left(M\right). ∎

In the remainder of this paper we will only work on compact GG-spaces. Theorem 4.1 establishes that the open and closed GG-covering numbers coincide for compact spaces, so we will always denote it simply by covG(M)\text{cov}_{G}\left(M\right), even though most of the times we will produce closed GG-covers.

The next Theorem explains the relation between the GG-covering of a space and the chromatic number of a GG-Borsuk graph on the same space. We choose to state this relation via two separate inequalities, since for some of the later results, we will only require one or the other. Theorem 3.3 will then follow immediately.

Theorem 4.2.

Let GG be a finite group, and MM a compact metric GG-space s.t. GG acts on MM via Lipschitz maps. Then the following hold.

  1. 1.

    There exists a constant DD s.t. if XMX\subset M is an (ε/D)(\varepsilon/D)-net, we have

    covG(M)χ(𝒢G(X,ε)).\text{cov}_{G}\left(M\right)\leq\chi(\mathcal{G}_{G}\left(X,\varepsilon\right)).
  2. 2.

    There exists a constant ε0>0\varepsilon_{0}>0 s.t. for any XMX\subset M, and any 0<ε<ε00<\varepsilon<\varepsilon_{0} we have

    χ(𝒢G(X,ε))covG(M).\chi(\mathcal{G}_{G}\left(X,\varepsilon\right))\leq\text{cov}_{G}\left(M\right).
Proof.

Let k=covG(M)k=\text{cov}_{G}\left(M\right), so we can write M=F1FkM=F_{1}\cup\dots\cup F_{k}, where the FiF_{i}’s are a closed GG-cover. Using the same notation as in the proof of Theorem 4.1 item 3, there exists b,D>0b,D>0 such that

b𝕕(x,g(y))x,yFi,i=1,,k,gG{𝟙}, andb\leq\mathbbm{d}\left(x,g(y)\right)\;\forall x,y\in F_{i},\;i=1,\dots,k,\;g\in G\setminus\{\mathbbm{1}\},\text{ and}
𝕕(g(x),g(y))(D1)𝕕(x,y)x,yM,gG{𝟙}.\mathbbm{d}\left(g(x),g(y)\right)\leq(D-1)\mathbbm{d}\left(x,y\right)\;\forall x,y\in M,\;g\in G\setminus\{\mathbbm{1}\}.
  1. 1.

    Let XMX\subset M be a (ε/D)(\varepsilon/D)-net of MM. Suppose by way of contradiction that there is a proper (k1)(k-1)-coloring for 𝒢G(X,ε)\mathcal{G}_{G}\left(X,\varepsilon\right), and fix such a coloring. For each color i=1,,k1i=1,\dots,k-1 define

    Ui=vXv has color iint(B(v,εD)).U_{i}=\bigcup_{\begin{subarray}{c}v\in X\\ v\text{ has color }i\end{subarray}}\text{int}\left(B\left(v,\frac{\varepsilon}{D}\right)\right).

    Since XX is an (ε/D)(\varepsilon/D)-net of MM, the UiU_{i}’s are an open cover of MM. Suppose there is an index ii and a non-identity gGg\in G such that xUig(Ui)x\in U_{i}\cap g(U_{i}). Thus x=g(x)x=g(x^{\prime}) for some xUix^{\prime}\in U_{i}. By definition, there are vertices v,vXv,v^{\prime}\in X of color ii such that 𝕕(x,v)ε/D\mathbbm{d}\left(x,v\right)\leq\varepsilon/D and 𝕕(x,v)ε/D\mathbbm{d}\left(x^{\prime},v^{\prime}\right)\leq\varepsilon/D. But then

    𝕕(v,g(v))\displaystyle\mathbbm{d}\left(v,g(v^{\prime})\right) 𝕕(v,x)+𝕕(x,g(v))=𝕕(v,x)+𝕕(g(x),g(v))\displaystyle\leq\mathbbm{d}\left(v,x\right)+\mathbbm{d}\left(x,g(v^{\prime})\right)=\mathbbm{d}\left(v,x\right)+\mathbbm{d}\left(g(x^{\prime}),g(v^{\prime})\right)
    𝕕(v,x)+(D1)𝕕(x,v)εD+(D1)εD=ε\displaystyle\leq\mathbbm{d}\left(v,x\right)+(D-1)\mathbbm{d}\left(x^{\prime},v^{\prime}\right)\leq\frac{\varepsilon}{D}+(D-1)\frac{\varepsilon}{D}=\varepsilon

    So vvv\sim v^{\prime} are adjacent in 𝒢G(X,ε)\mathcal{G}_{G}\left(X,\varepsilon\right), and they both have the same color ii, which is absurd. Hence Uig(Ui)=U_{i}\cap g(U_{i})=\emptyset for all ii and all gG{𝟙}g\in G\setminus\{\mathbbm{1}\}. But this is a (k1)(k-1) open cover of MM, which is a contradiction since covG(M)=k\text{cov}_{G}\left(M\right)=k. Therefore kχ(𝒢G(X,ε))k\leq\chi(\mathcal{G}_{G}\left(X,\varepsilon\right)).

  2. 2.

    Define ε0=b\varepsilon_{0}=b, and let 0<ε<ε00<\varepsilon<\varepsilon_{0}. Let XMX\subset M. We can kk-color the graph 𝒢G(X,ε)\mathcal{G}_{G}\left(X,\varepsilon\right) via the coloring c:X{1,,k}c:X\to\{1,\dots,k\} given by c(v)=min{i:vFi}c(v)=\min\{i:v\in F_{i}\}. Indeed, if v,wXv,w\in X have c(v)=c(w)=ic(v)=c(w)=i, then v,wFiv,w\in F_{i}, so 𝕕(v,g(w))b>ε\mathbbm{d}\left(v,g(w)\right)\geq b>\varepsilon for all gG{𝟙}g\in G\setminus\{\mathbbm{1}\}. Hence v≁wv\not\sim w, and cc is a proper kk-coloring. Therefore χ(𝒢G(X,ε))k\chi(\mathcal{G}_{G}\left(X,\varepsilon\right))\leq k.∎

We finish this section on general properties, by computing the clique number of GG-Borsuk graphs. In a sense, this can be seen as a generalization of Lemma 2.2 in [KMF20], regarding the odd-girth of 2\mathbb{Z}_{2}-Borsuk graphs.

Lemma 4.3.

Let MM be a compact GG-space such that GG acts via Lipschitz maps. Then, there exist constants ε0\varepsilon_{0} and DD such that, whenever XMX\subset M is a (ε/D)(\varepsilon/D)-net of XX and ε<ε0\varepsilon<\varepsilon_{0}, then ω(𝒢G(X,ε))=|G|\omega\left(\mathcal{G}_{G}\left(X,\varepsilon\right)\right)=|G|.

Proof.

Let b=min{𝕕(x,g(x)):xM,gG,g𝟙}b=\min\{\mathbbm{d}\left(x,g(x)\right):x\in M,g\in G,g\neq\mathbbm{1}\}. Note b>0b>0 since the GG-action is free and MM is compact. Let m=|G|m=|G| and enumerate G={g1=𝟙,g2,,gm}G=\{g_{1}=\mathbbm{1},g_{2},\dots,g_{m}\}. Let λ\lambda be the largest Lipschitz constant of the maps gig_{i}. Then set D=2max{λ,λ2}+2D=2\max\{\lambda,\lambda^{2}\}+2 and ε0=b/D\varepsilon_{0}=b/D. Suppose now that XMX\subset M is an (ε/D)(\varepsilon/D)-net with 0<ε<ε00<\varepsilon<\varepsilon_{0}. Let zMz\in M be any point, thus for each i=1,,mi=1,\dots,m, there is a point xiXx_{i}\in X such that 𝕕(xi,giz)ε/D\mathbbm{d}\left(x_{i},g_{i}z\right)\leq\varepsilon/D. Then, for any iji\neq j we have

𝕕(xi,(gigj1)xj)\displaystyle\mathbbm{d}\left(x_{i},(g_{i}g_{j}^{-1})x_{j}\right) λ𝕕(gi1xi,gj1xj)λ(𝕕(gi1xi,z)+𝕕(gj1xj,z))\displaystyle\leq\lambda\mathbbm{d}\left(g_{i}^{-1}x_{i},g_{j}^{-1}x_{j}\right)\leq\lambda\left(\mathbbm{d}\left(g_{i}^{-1}x_{i},z\right)+\mathbbm{d}\left(g_{j}^{-1}x_{j},z\right)\right)
λ2(𝕕(xi,gixi)+𝕕(xj,gjz))λ2εD<ε.\displaystyle\leq\lambda^{2}\left(\mathbbm{d}\left(x_{i},g_{i}x_{i}\right)+\mathbbm{d}\left(x_{j},g_{j}z\right)\right)\leq\frac{\lambda^{2}\varepsilon}{D}<\varepsilon.

Since gigj1𝟙g_{i}g_{j}^{-1}\neq\mathbbm{1} whenever iji\neq j, xixjx_{i}\sim x_{j}, and so {x1,,xm}\{x_{1},\dots,x_{m}\} is a mm-clique in 𝒢G(X,ε)\mathcal{G}_{G}\left(X,\varepsilon\right).

On the other hand, suppose by way of contradiction that x0,,xmXx_{0},\dots,x_{m}\in X is a (m+1)(m+1)-clique. So, in particular, for each 1im1\leq i\leq m, there is an index ki1k_{i}\neq 1 s.t. 𝕕(x0,gkixi)ε\mathbbm{d}\left(x_{0},g_{k_{i}}x_{i}\right)\leq\varepsilon. Hence there must be two different values iji\neq j, such that gki=gkjg_{k_{i}}=g_{k_{j}}; fix these ii and jj. Since xixjx_{i}\sim x_{j}, there must also be an element gG,g𝟙g\in G,g\neq\mathbbm{1} such that 𝕕(xj,gxi)ε\mathbbm{d}\left(x_{j},gx_{i}\right)\leq\varepsilon. Then, setting y=gki1x0=gkj1x0y=g_{k_{i}}^{-1}x_{0}=g_{k_{j}}^{-1}x_{0}, we get

𝕕(xi,gxi)\displaystyle\mathbbm{d}\left(x_{i},gx_{i}\right) 𝕕(xi,y)+𝕕(xj,y)+𝕕(xj,gxi)\displaystyle\leq\mathbbm{d}\left(x_{i},y\right)+\mathbbm{d}\left(x_{j},y\right)+\mathbbm{d}\left(x_{j},gx_{i}\right)
=𝕕(xi,gki1x0)+𝕕(xj,gkj1x0)+𝕕(xj+gxi)\displaystyle=\mathbbm{d}\left(x_{i},g_{k_{i}}^{-1}x_{0}\right)+\mathbbm{d}\left(x_{j},g_{k_{j}}^{-1}x_{0}\right)+\mathbbm{d}\left(x_{j}+gx_{i}\right)
λ𝕕(gkixi,x0)+λ𝕕(gkjxj,x0)+ε(2λ+1)ε(2λ+1)ε0<b\displaystyle\leq\lambda\mathbbm{d}\left(g_{k_{i}}x_{i},x_{0}\right)+\lambda\mathbbm{d}\left(g_{k_{j}}x_{j},x_{0}\right)+\varepsilon\leq(2\lambda+1)\varepsilon\leq(2\lambda+1)\varepsilon_{0}<b

But this is a contradiction, so there are not (m+1)(m+1)-cliques, and ω(χ(𝒢G(X,ε)))=m=|G|\omega(\chi(\mathcal{G}_{G}\left(X,\varepsilon\right)))=m=|G|. ∎

5 Lower Bound

In this section we focus on proving the lower bound of Theorem 3.4. This result is topological, by bounding the GG-index of the cellular structure of Hom(Km,H)\text{Hom}(K_{m},H), following ideas of [Lov78], [BK03] and [BK06]. It is worth noting, that similar GG-actions on Hom-Posets have been independently studied in [Dan18]. We start by restating the result we will prove.

Theorem 5.1.

Let GG be a finite group and KK a geometric GG-simplicial complex. Let k=indG(K)k=\text{ind}_{G}\left(K\right). Then k+|G|covG(K)k+|G|\leq\text{cov}_{G}\left(K\right).

Given two graphs H1H_{1} and H2H_{2}, the set of graph homomorphisms Hom(H1,H2)\text{Hom}(H_{1},H_{2}) can be endowed with a topology as a prodsimplicial complex, as introduced by Babson and Kozlov in [BK03]. Let us explain this structure.

Definition 5.2.

A cell complex is called prodsimplicial if all its cells are products of simplices. I.e. if all of its cells are of the form σ=σ1××σk\sigma=\sigma_{1}\times\dots\times\sigma_{k}, where σi\sigma_{i} is a did_{i}-simplex. Notation instead of using the ``×′′``\times^{\prime\prime} symbol, we will write σ=(σ1,,σk)\sigma=(\sigma_{1},\dots,\sigma_{k}).

Definition 5.3.

Given H1H_{1} and H2H_{2}, Hom(H1,H2)\text{Hom}(H_{1},H_{2}) is a prodsimplicial complex with cells σ\sigma satisfying:

  1. 1.

    σ=xV(H1)σx\sigma=\prod_{x\in V(H_{1})}\sigma_{x}, where σxV(H2)\sigma_{x}\subset V(H_{2}). That is, σ\sigma is the direct product of simplices on the vertices of H2H_{2}, indexed by the vertices of H1H_{1}, and

  2. 2.

    If φ:V(H1)V(H2)\varphi:V(H_{1})\to V(H_{2}) is such that φ(x)σx\varphi(x)\in\sigma_{x} for all xV(H2)x\in V(H_{2}), then it extends to a graph homomorphism φ:H1H2\varphi:H_{1}\to H_{2}.

We highlight the following points:

  • The cells of Hom(H1,H2)\text{Hom}(H_{1},H_{2}) can be though of as indexed by functions η:V(H1)(2V(H2))\eta:V(H_{1})\to\left(2^{V(H_{2})}\setminus\emptyset\right), such that if {x,y}E(H1)\{x,y\}\in E(H_{1}) then η(x)×η(y)E(H2)\eta(x)\times\eta(y)\subset E(H_{2}).

  • The dimension of the cell σ\sigma is given by xV(G)dim(σx)\sum_{x\in V(G)}\text{dim}(\sigma_{x}). Thus the vertices (0-cells) of Hom(H1,H2)\text{Hom}(H_{1},H_{2}) are precisely the graph homomorphisms η:GH\eta:G\to H.

  • The 1-skeleton can be seen as a graph, with vertices the graph homomorphisms and edges φψ\varphi\sim\psi if and only if they differ at exactly one value.

Lemma 5.4.

Let KmK_{m} denote the complete graph with mm vertices. If mtm\leq t, then

dim(Hom(Km,Kt))=tm\dim\left(\text{Hom}(K_{m},K_{t})\right)=t-m
Proof.

Let σ\sigma be a cell of Hom(Km,Kt)\text{Hom}(K_{m},K_{t}). Thus σ=i=1mσi=(σ1,,σm)\sigma=\prod_{i=1}^{m}\sigma_{i}=(\sigma_{1},\dots,\sigma_{m}) where each σi{1,,t}\sigma_{i}\subset\{1,\dots,t\}. By definition, when iji\neq j, for all possible choices viσiv_{i}\in\sigma_{i} and vjσjv_{j}\in\sigma_{j}, we must have vivjv_{i}\sim v_{j} in KtK_{t}. This is true iff vivjv_{i}\neq v_{j}. Hence, σiσj=\sigma_{i}\cap\sigma_{j}=\emptyset for iji\neq j. Then

dim(σ)=i=1mdim(σi)=i=1m(|σi|1)=(i=1m|σi|)m=|i=1mσi|m|{1,,t}|m=tm.\dim(\sigma)=\sum_{i=1}^{m}\dim(\sigma_{i})=\sum_{i=1}^{m}(|\sigma_{i}|-1)=\left(\sum_{i=1}^{m}|\sigma_{i}|\right)-m=\left|\bigcup_{i=1}^{m}\sigma_{i}\right|-m\leq|\{1,\dots,t\}|-m=t-m.

Moreover, the dimension is exactly tmt-m, since σ=(1,2,3,,m1,{m,m+1,,t})\sigma=(1,2,3,\dots,m-1,\{m,m+1,\dots,t\}) is a (tm)(t-m)-dimensional cell. ∎

5.1 Hom(𝑲𝒎,)\bm{\text{Hom}(K_{m},\cdot)} as 𝑮\bm{G}-space.

Let GG be a finite group with |G|=m|G|=m. Enumerate G={g1=𝟙,g2,g3,,gm}G=\{g_{1}=\mathbbm{1},g_{2},g_{3},\dots,g_{m}\}. By identifying the elements of GG with the vertices {1,,m}\{1,\dots,m\} of the complete graph KmK_{m}, GG acts on KmK_{m} via right multiplication. For a finite graph HH, GG acts on Hom(Km,H)\text{Hom}(K_{m},H) as follows: If gGg\in G, and σ=(σ1,,σm)\sigma=(\sigma_{1},\dots,\sigma_{m}), then gσ=(σπ(1),,σπ(m))g\sigma=(\sigma_{\pi(1)},\dots,\sigma_{\pi(m)}), where π(i)\pi(i) is the unique index such that gig=gπ(i)g_{i}g=g_{\pi(i)}. Note that if g𝟙g\neq\mathbbm{1}, then π(i)i\pi(i)\neq i for all ii. As we pointed out in the proof of Lemma 5.4, if σ\sigma is a cell in Hom(Km,H)\text{Hom}(K_{m},H), it must be that σiσj=\sigma_{i}\cap\sigma_{j}=\emptyset for all iji\neq j, in particular,

σg(σ)(σ1σπ(1),,σmσπ(m))=.\sigma\cap g(\sigma)\subset(\sigma_{1}\cap\sigma_{\pi(1)},\dots,\sigma_{m}\cap\sigma_{\pi(m)})=\emptyset.

So GG acts freely on Hom(G,H)\text{Hom}(G,H) via cellular maps. Moreover, it is clear that this construction is functorial, so if f:HHf:H\to H^{\prime} is a graph homomorphism, then it induces a GG-map

f:Hom(Km,H)𝐺Hom(Km,H).f:\text{Hom}(K_{m},H)\xrightarrow{G}\text{Hom}(K_{m},H^{\prime}).

5.2 Proof of Theorem 5.1

Let KK be a free GG-simplicial complex with k=indG(K)k=\text{ind}_{G}\left(K\right) and m=|G|m=|G|. Fix ε>0\varepsilon>0. Let TT be a finer triangulation of KK, such that TT is also a GG-simplicial complex and diam(σ)ε\text{diam}\left(\sigma\right)\leq\varepsilon for all faces σT\sigma\in T. Let HH be the graph with vertex set V(T)V(T) and edges xyx\sim y whenever there exists gG{𝟙}g\in G\setminus\{\mathbbm{1}\} such that {x,g(y)}T(1)\{x,g(y)\}\in T^{(1)}. Note then HH is an induced subgraph of the GG-Borsuk graph 𝒢G(V(T),ε)\mathcal{G}_{G}\left(V(T),\varepsilon\right).

Thus if ε<ε0\varepsilon<\varepsilon_{0}, for ε0\varepsilon_{0} the constant given by Theorem 4.2 item 2, we get

χ(H)χ(𝒢G(V(T),ε))covG(K).\chi(H)\leq\chi(\mathcal{G}_{G}\left(V(T),\varepsilon\right))\leq\text{cov}_{G}\left(K\right).

We will construct a GG-map Φ:T𝐺Hom(Km,H)\Phi:T\xrightarrow{G}\text{Hom}(K_{m},H). Since HKχ(H)H\to K_{\chi(H)}, the functoriality of the GG-action on Hom(Km,)\text{Hom}(K_{m},\cdot) produces

T𝐺Hom(Km,H)𝐺Hom(Km,Kχ(H)).T\xrightarrow{G}\text{Hom}(K_{m},H)\xrightarrow{G}\text{Hom}(K_{m},K_{\chi(H)}).

Finally, applying items 1 and 5 of Theorem 2.7 and the dimension computed in Lemma 5.4, we get

k\displaystyle k =indG(K)=indG(T)indG(Hom(Km,H))\displaystyle=\text{ind}_{G}\left(K\right)=\text{ind}_{G}\left(T\right)\leq\text{ind}_{G}\left(\text{Hom}(K_{m},H)\right)
indG(Hom(Km,Kχ(H))dim(Hom(Km,Kχ(H)))χ(H)m,\displaystyle\leq\text{ind}_{G}\left(\text{Hom}(K_{m},K_{\chi(H)}\right)\leq\dim(\text{Hom}(K_{m},K_{\chi(H)}))\leq\chi(H)-m,

so k+mχ(H)covG(K)k+m\leq\chi(H)\leq\text{cov}_{G}\left(K\right) as desired. From now on, we focus on defining the map Φ\Phi.

Enumerate G={g1=𝟙,g2,g3,,gm}G=\{g_{1}=\mathbbm{1},g_{2},g_{3},\dots,g_{m}\}. Let σ\sigma be a dd-dimensional face of KK, for any 0ddim(K)0\leq d\leq\dim(K). Let Ψσ=gGg(σ)=(σ,g2(σ),g3(σ),,gm(σ))\Psi_{\sigma}=\prod_{g\in G}g(\sigma)=(\sigma,g_{2}(\sigma),g_{3}(\sigma),\dots,g_{m}(\sigma)). Note Ψσ\Psi_{\sigma} is isomorphic to the contractible prodsimplicial complex (Δd)m(\Delta_{d})^{m}, where Δd\Delta_{d} is the dd-dimensional simplex, and the power mm refers to the Cartesian Product.

We claim Ψσ\Psi_{\sigma} is a subcomplex of Hom(Km,H)\text{Hom}(K_{m},H). Indeed gi(σ)V(H)g_{i}(\sigma)\subset V(H), so Ψσ\Psi_{\sigma} is the product of simplices in V(H)V(H) indexed by the vertices of KmK_{m}. To prove our claim, we need that whenever iji\sim j in KmK_{m}, i.e. iji\neq j, any choice xgi(σ)x\in g_{i}(\sigma) and ygj(σ)y\in g_{j}(\sigma) produces an edge xyx\sim y in HH. Pick any such xx and yy, so x=gi(v)x=g_{i}(v) and y=gj(w)y=g_{j}(w) for some vertices v,wσv,w\in\sigma. Put g=gigj1g=g_{i}g_{j}^{-1}, then

{x,g(y)}={gi(v),gi(w)}=gi({v,w})\{x,g(y)\}=\{g_{i}(v),g_{i}(w)\}=g_{i}(\{v,w\})

Since {v,w}σ\{v,w\}\subset\sigma is an edge in TT, so is gi({v,w})g_{i}(\{v,w\}), and thus xyx\sim y in HH as needed. Moreover, note that for any gGg\in G,

g(Ψσ)\displaystyle g(\Psi_{\sigma}) =g(g1(σ),g2(σ),,gm(σ))\displaystyle=g(g_{1}(\sigma),g_{2}(\sigma),\dots,g_{m}(\sigma))
=(gπ(1)(σ),gπ(2)(σ),,gπ(m)(σ))\displaystyle=(g_{\pi(1)}(\sigma),g_{\pi(2)}(\sigma),\dots,g_{\pi(m)}(\sigma))
=(g1g(σ),g2g(σ),,gmg(σ))=Ψg(σ).\displaystyle=(g_{1}g(\sigma),g_{2}g(\sigma),\dots,g_{m}g(\sigma))=\Psi_{g(\sigma)}.

Note also, if τ\tau is a face of σ\sigma, Ψτ\Psi_{\tau} is a subcomplex of Ψσ\Psi_{\sigma}, since for any gGg\in G, g(τ)g(σ)g(\tau)\subset g(\sigma).

Recall we say that a topological space XX is nn-connected if for every =1,0,,n\ell=-1,0,\dots,n, each continuous map f:𝕊Xf:\mathbb{S}^{\ell}\to X can be extended to a continuous map f:B+1Xf:B^{\ell+1}\to X. It is well known that any contractible space is nn-connected for all n0n\geq 0. Recall the dd-simplex Δd\Delta_{d} is homeomorphic to the closed ball Bd+1B^{d+1} and its boundary Δd\partial\Delta_{d} is homeomorphic to the (d1)(d-1)-sphere 𝕊d1\mathbb{S}^{d-1}. Putting these together with the fact Ψσ\Psi_{\sigma} is contractible, we get the following tool:

If f:σΨσf:\partial\sigma\to\|\Psi_{\sigma}\| is a continuous map, then it can be extended into a map f:σΨσf:\sigma\to\|\Psi_{\sigma}\|.

We now construct Φ\Phi by induction on the dd-skeleton of TT.

  • Vertices: For each vertex vV(T)v\in V(T), define Φ(v)=(v,g2(v),,gm(v))=Ψ{v}\Phi(v)=(v,g_{2}(v),\dots,g_{m}(v))=\Psi_{\{v\}}. Clearly Φ(g(v))=Ψ{g(v)}=g(Ψ{v})=g(Φ(v))\Phi(g(v))=\Psi_{\{g(v)\}}=g(\Psi_{\{v\}})=g(\Phi(v)) as proved above.

  • Induction Hypothesis: Suppose that we have defined Φ:T(d)𝐺Hom(Km,H)\Phi:T^{(d)}\xrightarrow{G}\text{Hom}(K_{m},H) for the dd-skeleton of TT, in such a way that Φ(σ)Ψσ\Phi(\sigma)\subset\|\Psi_{\sigma}\|.

  • Inductive Step: Separate the (d+1)(d+1)-cells of TT into GG-orbits (this is possible because GG acts freely). Let 𝒪=Gσ\mathcal{O}=G\sigma be one of such orbits, with σT(d+1)\sigma\in T^{(d+1)} a generator. Let τ0,τ1,,τd\tau_{0},\tau_{1},\dots,\tau_{d} be the dd-faces of σ\sigma. By the induction hypothesis, Φ\Phi is defined such that Φ(τi)ΨτiΨσ\Phi(\tau_{i})\subset\|\Psi_{\tau_{i}}\|\subset\|\Psi_{\sigma}\|. Thus Φ\Phi is defined as a map

    Φ:σ=i=0dτii=0dΨτiΨσ,\Phi:\partial\sigma=\bigcup_{i=0}^{d}\tau_{i}\to\bigcup_{i=0}^{d}\|\Psi_{\tau_{i}}\|\subset\|\Psi_{\sigma}\|,

    hence we can extend Φ\Phi to σ\sigma, such that Φ(σ)Ψσ\Phi(\sigma)\subset\|\Psi_{\sigma}\|. For each other (d+1)(d+1)-cell ω𝒪\omega\in\mathcal{O}, there is a gGg\in G, such that ω=g(σ)\omega=g(\sigma), so define

    Φ(ω):=g(Φ(σ))g(Ψσ)Ψgσ=Ψω.\Phi(\omega):=g(\Phi(\sigma))\subset g\left(\|\Psi_{\sigma}\|\right)\subset\|\Psi_{g\sigma}\|=\|\Psi_{\omega}\|.

    Note this agrees on the dd-faces of ω\omega, since we supposed Φ\Phi was a GG-map on the dd-skeleton of TT. Finally, repeating this for all the GG-orbits of (d+1)(d+1)-cells, we get Φ:T(d+1)𝐺Hom(Km,H)\Phi:T^{(d+1)}\xrightarrow{G}\text{Hom}(K_{m},H) satisfying the induction hypothesis.

This produces the GG-map Φ:T𝐺Hom(Km,H)\Phi:T\xrightarrow{G}\text{Hom}(K_{m},H) and finishes the proof.

6 Upper Bounds

We now focus on the upper bound in Theorem 3.4. Note that if KK is a GG-space of GG-index kk, K𝐺EkGK\xrightarrow{G}E_{k}G, so by Theorem 4.1, covG(K)covG(EkG)=covG(k)\text{cov}_{G}\left(K\right)\leq\text{cov}_{G}\left(E_{k}G\right)=\text{cov}_{G}\left(k\right). Hence, we only need to prove an upper bound for finite GG classifying spaces. We do this by combining Theorem 6.1, which provides a recursive upper bound, and Theorem 6.2, which bounds covG(1)\text{cov}_{G}\left(1\right). We finish this section with Theorem 6.5, which we use to compute upper bounds for covm(d)\text{cov}_{\mathbb{Z}_{m}}\left(d\right), for some values of mm and dd, with the aid of a computer.

Theorem 6.1.

Let GG be a finite group and d2d\geq 2 an integer. Then

covG(d)covG(d1)+|G|1.\text{cov}_{G}\left(d\right)\leq\text{cov}_{G}\left(d-1\right)+|G|-1.
Proof.

Let K=GMK=G*M where MM is a (d1)(d-1)-dimensional classifying space for GG, so KK is a dd-dimensional classifying space for GG. The elements of KK can be described as linear combinations tg(1t)μtg\oplus(1-t)\mu where gG,μMg\in G,\mu\in M and 0t10\leq t\leq 1, and thus hGh\in G acts by

h(tg(1t)μ)=hg(1t)h(μ).h(tg\oplus(1-t)\mu)=hg\oplus(1-t)h(\mu).

Let k=covG(d1)k=\text{cov}_{G}\left(d-1\right), so there is a closed GG-cover M=F1,,FkM=F_{1},\dots,F_{k}. For each i=1,,ki=1,\dots,k, and each gG,g𝟙g\in G,g\neq\mathbbm{1}, define the closed sets

Fi,g\displaystyle F_{i,g} ={tg(1t)μK:μFi,0t1/2} and\displaystyle=\left\{tg\oplus(1-t)\mu\in K:\mu\in F_{i},0\leq t\leq 1/2\right\}\text{ and }
F¯i,g\displaystyle\overline{F}_{i,g} ={tg(1t)μK:μFi,1/2t1}.\displaystyle=\left\{tg\oplus(1-t)\mu\in K:\mu\in F_{i},1/2\leq t\leq 1\right\}.

So Fi,gF¯i,g=gFiF_{i,g}\cup\overline{F}_{i,g}=g*F_{i}. Also define the closed sets

Fi,𝟙\displaystyle F_{i,\mathbbm{1}} ={t𝟙(1t)μK:μFi,0t1}=𝟙Fi.\displaystyle=\left\{t\mathbbm{1}\oplus(1-t)\mu\in K:\mu\in F_{i},0\leq t\leq 1\right\}=\mathbbm{1}*F_{i}.

Finally, let m=|G|m=|G| and enumerate G={𝟙,g1,g2,,gm1}G=\{\mathbbm{1},g_{1},g_{2},\dots,g_{m-1}\}. Define E1,,Ek,Ek+1,,Em+k1E_{1},\dots,E_{k},E_{k+1},\dots,E_{m+k-1} as follows

Ei\displaystyle E_{i} =gGFi,g for 1ik and\displaystyle=\bigcup_{g\in G}F_{i,g}\text{ for }1\leq i\leq k\text{ and}
Ek+i\displaystyle E_{k+i} =i=1kF¯j,gi for 1im1.\displaystyle=\bigcup_{i=1}^{k}\overline{F}_{j,g_{i}}\text{ for }1\leq i\leq m-1.

Clearly the (m+k1)(m+k-1) sets EiE_{i} are closed and cover KK. We claim they are a GG-cover of KK.

Suppose there is some index 1ik1\leq i\leq k and some hG,h𝟙h\in G,h\neq\mathbbm{1} such that EihEiE_{i}\cap hE_{i}\neq\emptyset. Then, there must be some g,gGg,g^{\prime}\in G such that xFi,ghFi,gx\in F_{i,g}\cap hF_{i,g^{\prime}}\neq\emptyset. Thus we can write xx as

x=tg(1t)μ=τhg(1τ)hν,x=tg\oplus(1-t)\mu=\tau hg^{\prime}\oplus(1-\tau)h\nu,

for some μ,νFi\mu,\nu\in F_{i}, 0t,τ10\leq t,\tau\leq 1. Then, it must be that (1t)μ=(1τ)hν(1-t)\mu=(1-\tau)h\nu. If (1t),(1τ)0(1-t),(1-\tau)\neq 0 we must have μ=hν\mu=h\nu, but since FihFi=F_{i}\cap hF_{i}=\emptyset, this is impossible. If instead (1t)=(1τ)=0(1-t)=(1-\tau)=0, then it has to be that g=g=𝟙g=g^{\prime}=\mathbbm{1} and then x=𝟙=h𝟙=hx=\mathbbm{1}=h\mathbbm{1}=h, but h𝟙h\neq\mathbbm{1} so this is impossible as well.

Suppose there is some index 1im11\leq i\leq m-1, and some hG,h𝟙h\in G,h\neq\mathbbm{1} such that Ek+ihEk+iE_{k+i}\cap hE_{k+i}\neq\emptyset. Then, there must be some 1j,αk1\leq j,\alpha\leq k such that xF¯j,gihF¯α,gix\in\overline{F}_{j,g_{i}}\cap h\overline{F}_{\alpha,g_{i}}\neq\emptyset. Thus we can write xx as

x=tgi(1t)μ=τhgi(1τ)hν,x=tg_{i}\oplus(1-t)\mu=\tau hg_{i}\oplus(1-\tau)h\nu,

for some μFj\mu\in F_{j}, νFα\nu\in F_{\alpha}, 1/2t,τ11/2\leq t,\tau\leq 1. Then, it must be that tgi=τhgitg_{i}=\tau hg_{i}, and since t,τ0t,\tau\neq 0, we must have gi=hgig_{i}=hg_{i}, but since h𝟙h\neq\mathbbm{1}, this is impossible.

Therefore ErhEr=E_{r}\cap hE_{r}=\emptyset for all 1rk+m11\leq r\leq k+m-1 and all hG,h𝟙h\in G,h\neq\mathbbm{1} as desired. ∎

Now we state the upper bound for the 1-dimensional case.

Theorem 6.2.

Let GG a finite group. Then

covG(1)|G|+1.\text{cov}_{G}\left(1\right)\leq|G|+1.

Combining this Theorem with the lower bound in Theorem 5.1, gives covG(1)=|G|+1\text{cov}_{G}\left(1\right)=|G|+1. To prove it, we need to first establish some lemmas. Let us start proving the case when GG is a cyclic group acting on the 1-dimensional classifying space 𝕊1\mathbb{S}^{1}.

Lemma 6.3.

Consider 𝕊1\mathbb{S}^{1} as a m\mathbb{Z}_{m} space, then

covm(𝕊1)m+1.\text{cov}_{\mathbb{Z}_{m}}\left(\mathbb{S}^{1}\right)\leq m+1.
Proof.

Note m\mathbb{Z}_{m} acts on 𝕊1\mathbb{S}^{1} via rotations of 2π/m2\pi/m. Enumerate m={𝟙,ν,,νm1}\mathbb{Z}_{m}=\{\mathbbm{1},\nu,\dots,\nu^{m-1}\}. Note that for any x𝕊1x\in\mathbb{S}^{1} and any 1km11\leq k\leq m-1 we have

𝕕𝕊1(x,νk(x))𝕕𝕊1(x,ν(x))=2πm.\mathbbm{d}_{\mathbb{S}^{1}}\left(x,\nu^{k}(x)\right)\geq\mathbbm{d}_{\mathbb{S}^{1}}\left(x,\nu(x)\right)=\frac{2\pi}{m}.

For each i=0,,mi=0,\dots,m, let FiF_{i} be the closed circular arc [2iπm+1,2(i+1)πm+1][\frac{2i\pi}{m+1},\frac{2(i+1)\pi}{m+1}]. Clearly diam(Fi)=2πm+1<2πm\text{diam}\left(F_{i}\right)=\frac{2\pi}{m+1}<\frac{2\pi}{m}, so it is impossible to have xx and νk(x)\nu^{k}(x) on the same FiF_{i}, that is FiνkFi=F_{i}\cap\nu^{k}F_{i}=\emptyset for all i=0,,mi=0,\dots,m and all k=1,,m1k=1,\dots,m-1. Therefore, the FiF_{i}’s are a m\mathbb{Z}_{m}-cover of 𝕊1\mathbb{S}^{1} using (m+1)(m+1) closed sets, and the result follows. ∎

Lemma 6.4.

Let K=mmK=\mathbb{Z}_{m}*\mathbb{Z}_{m}, so KK is trivially a m\mathbb{Z}_{m}-space. As a simplicial complex, the vertices of KK are two disjoint copies of m\mathbb{Z}_{m}. Denote these vertices by 1,,m1,\dots,m if they come from the first copy, and 1¯,,m¯\overline{1},\dots,\overline{m}, from the second. Then, there exists a closed m\mathbb{Z}_{m}-cover K=A0AmK=A_{0}\cup\dots\cup A_{m}, with (m+1)(m+1) closed sets, such that i,i¯Aii,\overline{i}\in A_{i} for each im={1,,m}i\in\mathbb{Z}_{m}=\{1,\dots,m\}.

Proof.

Let Φ:K𝕊1\Phi:K\to\mathbb{S}^{1} be defined as follows. For the vertices of KK, let Φ(i)=Φ(i¯)=(2πi)/m𝕊1\Phi(i)=\Phi(\overline{i})=(2\pi i)/m\in\mathbb{S}^{1}. For the edges {i,j¯}K\{i,\overline{j}\}\in K, Φ({i,j¯})=[2πim,2πjk]\Phi(\{i,\overline{j}\})=[\frac{2\pi i}{m},\frac{2\pi j}{k}]. This defines Φ\Phi on the simplicial complex KK, and is straightforward to check it is both continuous and a m\mathbb{Z}_{m}-map.

Lemma 6.3 ensures there is a m\mathbb{Z}_{m}-cover F0,,FmF_{0},\dots,F_{m} of 𝕊1\mathbb{S}^{1}. Clearly the points 2πim\frac{2\pi i}{m} must be in different sets FiF_{i} for it to be a m\mathbb{Z}_{m}-cover, so we may assume 2πimFi\frac{2\pi i}{m}\in F_{i} for i=1,,mi=1,\dots,m. Finally, the pullback cover Ai=Φ1(Fi)A_{i}=\Phi^{-1}(F_{i}) for i=0,,mi=0,\dots,m produces the desired m\mathbb{Z}_{m}-cover. ∎

Regarding the closed sets of a m\mathbb{Z}_{m}-cover as colors, we can give an intuitive interpretation of Lemma 6.4. If the vertices of KK already have colors assigned, such that c(i)=c(i¯)c(i)=c(\overline{i}) for all ii, and c(i)c(j)c(i)\neq c(j) whenever iji\neq j, then we can extend such coloring to KK, adding just one more color in order to get a m\mathbb{Z}_{m}-cover. This interpretation is the key to prove Theorem 6.2, where finding a GG-cover for GGG*G can be broken into finding GG-covers for disjoint cyclic orbits of edges, using only the colors already assigned to the vertices and at most one more color.

Proof of Theorem 6.2.

Let m=|G|m=|G| and enumerate G={g1=𝟙,g2,,gm}G=\{g_{1}=\mathbbm{1},g_{2},\dots,g_{m}\}. Let K=GGK=G*G, and denote its vertices by g1,,gmg_{1},\dots,g_{m} and g1¯,,gm¯\overline{g_{1}},\dots,\overline{g_{m}}, as in Lemma 6.4. All points xKx\in\|K\| can be written as tgi(1t)gj¯tg_{i}\oplus(1-t)\overline{g_{j}}, for some gi,gjGg_{i},g_{j}\in G and some 0t10\leq t\leq 1. The edges of KK are all edges of the form {gi,gj¯}\{g_{i},\overline{g_{j}}\}, for 1i,jm1\leq i,j\leq m. To ease our notation, we will denote an edge {gi,gj¯}\{g_{i},\overline{g_{j}}\} simply as gigjg_{i}\sim g_{j}, where it is understood that the element on the left denotes that vertex and the element on the right denotes the overlined version of the vertex. In particular, ggg\sim g is valid, and denotes the edge {g,g¯}K\{g,\overline{g}\}\in K for any gGg\in G.

We will construct a closed GG-cover of KK by gradually adding closed subsets of KK to the initial color classes C0,CmC_{0},\dots C_{m}, initialized as C0=C_{0}=\emptyset and Ci={gi}{gi¯}C_{i}=\{g_{i}\}\cup\{\overline{g_{i}}\} for 1im1\leq i\leq m. Note they are indeed closed, since they are either empty or exactly two points.

For an easier exposition, we will say that a closed subset FKF\subset\|K\| is colored with color gig_{i}, to mean that we add the set FF to CiC_{i}, defining Ci=CiFiC_{i}^{\prime}=C_{i}\cup F_{i}. This is a slight abuse of notation, since it is possible for some points to be part of many sets CiC_{i}. Note that by doing this finitely many times, the color classes will remain closed.

We want that CigCi=C_{i}\cap gC_{i}=\emptyset for all gG,g𝟙g\in G,g\neq\mathbbm{1}, 0im0\leq i\leq m. This failing is equivalent to the existence of a point xKx\in\|K\|, an index 0im0\leq i\leq m and an element gGg\in G, g𝟙g\neq\mathbbm{1} such that xx and gxgx both have color gig_{i}. Since in our initial coloring the vertices of KK have different colors, xx must be in the interior of some edge aba\sim b, and thus gxgx is in the edge gagbga\sim gb. Thus, to prevent the existence of such xx, we just need to focus on properly extending the color classes C0,,CmC_{0},\dots,C_{m}, on the GG-orbit 𝒪(ab):={gagb:gG}\mathcal{O}(a\sim b):=\{ga\sim gb:g\in G\}, for each edge abKa\sim b\in K separately. When we say properly extending, we mean that C0𝒪(ab),,Cm𝒪(ab)C_{0}\cap\mathcal{O}(a\sim b),\dots,C_{m}\cap\mathcal{O}(a\sim b) is a closed GG-cover of 𝒪(ab)\mathcal{O}(a\sim b). We will now focus on doing this on one such orbit.

Let aba\sim b be an edge of KK, and 𝒪={gagb:gG}\mathcal{O}=\{ga\sim gb:g\in G\} be its GG-orbit. By interpreting each edge as an assignment σ(ga)=gb\sigma(ga)=gb, we can see σ\sigma as a permutation of the elements of GG. Recall that all finite permutations can be regarded as a disjoint union of cycles (see e.g. [Lan02, p. 30]), so σ\sigma partitions GG into rr cycles of the form

a1(1)σ(a1(1))=a2(1)σ(a2(1))\displaystyle a_{1}^{(1)}\mapsto\sigma(a_{1}^{(1)})=a_{2}^{(1)}\mapsto\sigma(a_{2}^{(1)}) =a3(1)σ(ak1(1))=a1(1)\displaystyle=a_{3}^{(1)}\mapsto\cdots\mapsto\sigma(a_{k_{1}}^{(1)})=a_{1}^{(1)}
a1(2)σ(a1(2))=a2(2)σ(a2(2))\displaystyle a_{1}^{(2)}\mapsto\sigma(a_{1}^{(2)})=a_{2}^{(2)}\mapsto\sigma(a_{2}^{(2)}) =a3(2)σ(ak2(2))=a1(2)\displaystyle=a_{3}^{(2)}\mapsto\cdots\mapsto\sigma(a_{k_{2}}^{(2)})=a_{1}^{(2)}
\displaystyle\vdots
a1(r)σ(a1(r))=a2(r)σ(a2(r))\displaystyle a_{1}^{(r)}\mapsto\sigma(a_{1}^{(r)})=a_{2}^{(r)}\mapsto\sigma(a_{2}^{(r)}) =a3(r)σ(akr(r))=a1(r)\displaystyle=a_{3}^{(r)}\mapsto\cdots\mapsto\sigma(a_{k_{r}}^{(r)})=a_{1}^{(r)}

where kjk_{j} is the length of the jj-th cycle.

And at the same time, this partitions 𝒪\mathcal{O} into rr disjoint subsets of the form

𝒪j={a1(j)a2(j),a2(j)a3(j),,akj(j)a1(j)}.\mathcal{O}_{j}=\left\{a_{1}^{(j)}\sim a_{2}^{(j)},a_{2}^{(j)}\sim a_{3}^{(j)},\dots,a_{k_{j}}^{(j)}\sim a_{1}^{(j)}\right\}.

Each 𝒪j\mathcal{O}_{j} can be seen itself as a 1-dimensional simplicial complex, and moreover it is naturally a kj\mathbb{Z}_{k_{j}}-space, where kj\mathbb{Z}_{k_{j}} acts by cyclically permuting the edges. Thus, we can regard 𝒪jkjkj\mathcal{O}_{j}\subset\mathbb{Z}_{k_{j}}*\mathbb{Z}_{k_{j}}. Lemma 6.4 gives a coloring for the whole space kjkj\mathbb{Z}_{k_{j}}*\mathbb{Z}_{k_{j}} just using the colors already assigned to the vertices ai(j)a_{i}^{(j)} and at most one more color (which we will assign to the color class C0C_{0}), so we may restrict this coloring to 𝒪j\mathcal{O}_{j}. In the especial case when σ\sigma is already a cycle, this finishes the proof. For the general case when r2r\geq 2, still Lemma 6.4 will be our building block, but we need to be careful so that the different cycles don’t produce points xx and gxgx with the same color.

Take each edge ai(j)ai+1(j)a_{i}^{(j)}\sim a_{i+1}^{(j)} (the addition i+1i+1 must be understood modulo kjk_{j}) and divide it into rr closed intervals Pi,1(j),Pi,2(j),,Pi,r(j)P_{i,1}^{(j)},P_{i,2}^{(j)},\dots,P_{i,r}^{(j)} intersecting only at their endpoints. More formally, we identify ai(j)ai+1(j)a_{i}^{(j)}\sim a_{i+1}^{(j)} isometrically with the real segment [0,1][0,1] by sending ai(j)a_{i}^{(j)} to 0 and ai+1(j)a_{i+1}^{(j)} to 1. Then Pk,i(j)P_{k,i}^{(j)} is the preimage of the subinterval [(k1)/n,k/n]\left[(k-1)/n,k/n\right]. Notice then that for any gGg\in G, and any 1kr1\leq k\leq r the set gPi,k(j)gP_{i,k}^{(j)} is always another interval Pi,k(j)P_{i^{\prime},k}^{(j^{\prime})}, i.e. it is part of a different edge that may be part of a different orbit 𝒪j\mathcal{O}_{j^{\prime}}, but will always be on the same relative position over the edge.

We finally produce the coloring for 𝒪={ai(j)ai+1(j):1ikj,1jr}\mathcal{O}=\{a_{i}^{(j)}\sim a_{i+1}^{(j)}:1\leq i\leq k_{j},1\leq j\leq r\} by coloring each subinterval Pi,k(j)P_{i,k}^{(j)}, as follows.

  1. 1.

    Color each subinterval Pi,k(j)P_{i,k}^{(j)} where k<jk<j with the color ai(j)a_{i}^{(j)}. I.e. color the first (j1)(j-1) subintervals of each edge, with the color of the left vertex.

  2. 2.

    Color each subinterval Pi,k(j)P_{i,k}^{(j)} where k>jk>j with the color ai+1(j)a_{i+1}^{(j)}. I.e. color the last (rj)(r-j) subintervals of each edge, with the color of the right vertex.

  3. 3.

    Fix each 1jr1\leq j\leq r. Notice that the endpoints of the subintervals Pi,j(j)P_{i,j}^{(j)}, were already assigned colors ai(j)a_{i}^{(j)} and ai+1(j)a_{i+1}^{(j)} respectively. By cycling on the edges ai(j)ai+1(j)a_{i}^{(j)}\sim a_{i+1}^{(j)}, we also cycle on the subintervals Pi,j(j)P_{i,j}^{(j)}, so themselves can be seen as a subset of the kj\mathbb{Z}_{k_{j}}-complex kjkj\mathbb{Z}_{k_{j}}*\mathbb{Z}_{k_{j}}. Thus lemma 6.4 gives a closed coloring for P1,j(j),,Pkj,j(j)P_{1,j}^{(j)},\dots,P_{k_{j},j}^{(j)} using only the colors already assigned to the ai(j)a_{i}^{(j)}’s and color 0 (color class C0C_{0}).

Checking that this coloring produces a closed GG-cover of 𝒪\mathcal{O}, is straightforward. By repeating this process for all the GG-orbits of edges in KK, we get the desired GG-cover with m+1m+1 closed sets. ∎

A possible proof of Conjecture 3.5, would be to adapt the technique of the proof of Theorem 6.2 to higher dimensions. Doing this, would require to first prove Conjecture 3.6 item 1, regarding m\mathbb{Z}_{m} spaces. While we haven’t been able to do this generally, even for dimension 2, we have been able to explicitly find some m\mathbb{Z}_{m}-covers for small values of mm aided by a computer. The following Theorem allows us to find GG-covers computing the chromatic number of relatively small graphs. More details on the specific graphs and techniques used to produce the examples of Figure 1(d) are included in the Appendix.

Theorem 6.5.

Let KK be a compact geometric GG-simplicial complex. Let TT be a triangulation of KK, such that it is still a GG-simplicial complex. Define the graph HH to be the graph with vertices V(T)V(T) and with edges vwv\sim w whenever there exists a gG,g𝟙g\in G,g\neq\mathbbm{1} such that {v,gw}T\{v,gw\}\in T. Then

covG(K)χ(H).\text{cov}_{G}\left(K\right)\leq\chi(H).
Proof.

If HH has loops, its chromatic number is infinite and the result is trivial; so suppose HH has no loops. Let cc be a coloring of HH using χ(H)\chi(H) colors, so c:V(T){1,,χ(H)}c:V(T)\to\{1,\dots,\chi(H)\}.

Let B=Bar(T)B=\text{Bar}(T) be the Barycentric subdivision of TT. Recall B=T=K\|B\|=\|T\|=\|K\| and that the simplices of BB can be identified with chains of simplices of TT. Since TT is a GG-simplicial complex, applying gg to a chain σ1σ2σr\sigma_{1}\prec\sigma_{2}\prec\dots\sigma_{r} of simplices of TT, produces gσ1gσ2gσrg\sigma_{1}\prec g\sigma_{2}\prec\dots\prec g\sigma_{r}, another chain of simplices of TT. So BB is a GG-simplicial complex as well. Recall that the facets of BB correspond to maximal chains of simplices of TT, so for each facet τB\tau\in B there is a unique vertex vV(T)v\in V(T) such that vτv\in\tau.

For each vertex vV(T)v\in V(T), we define F(v)F(v) as follows

F(v)={τ:τFacets(B) and vτ}F(v)=\bigcup\{\|\tau\|:\tau\in\text{Facets}(B)\text{ and }v\in\tau\}

Note that the geometric representations of the facets of BB give a closed cover of K\|K\|, and so the sets F(v)F(v) are also a closed cover. Also note that gF(v)=F(gv)gF(v)=F(gv) for all gG,vV(T)g\in G,v\in V(T).

Define the closed sets C1,,Cχ(H)C_{1},\dots,C_{\chi(H)} as

Ci=vV(T)c(v)=iF(v).C_{i}=\bigcup_{\begin{subarray}{c}v\in V(T)\\ c(v)=i\end{subarray}}F(v).

Then the sets CiC_{i} are a closed cover for K\|K\|, and we claim they are a GG-cover. To see this, suppose by way of contradiction that there is a point xx in K\|K\|, an index ii, and an element gGg\in G, g𝟙g\neq\mathbbm{1} such that x,gxCix,gx\in C_{i}. This means there are vertices u,vV(T)u,v\in V(T) with c(u)=c(v)=ic(u)=c(v)=i, such that xF(u)x\in F(u) and gxF(v)gx\in F(v). Thus gxgF(u)=F(gu)gx\in gF(u)=F(gu) and gxF(v)gx\in F(v).

Then, there are simplices τ1,τ2Facets(B)\tau_{1},\tau_{2}\in\text{Facets}(B), such that guτ1,vτ2gu\in\tau_{1},v\in\tau_{2} and gxτ1τ2gx\in\|\tau_{1}\|\cap\|\tau_{2}\|. Thus, there is a nontrivial face τ=τ1τ2\tau=\tau_{1}\cap\tau_{2} in BB. We claim this means {v,gu}T\{v,gu\}\in T, and so vuv\sim u in HH, but they both had the same color ii, giving the desired contradiction.

To finish the proof we establish now our last claim. Since guτ1gu\in\tau_{1}, vτ2v\in\tau_{2}, and τ=τ1τ2\tau=\tau_{1}\cap\tau_{2}\neq\emptyset, we can describe them as chains of simplices in TT:

τ1\displaystyle\tau_{1} ={gua2ar}\displaystyle=\{gu\prec a_{2}\prec\dots\prec a_{r}\}
τ2\displaystyle\tau_{2} ={vb2bs} and\displaystyle=\{v\prec b_{2}\prec\dots\prec b_{s}\}\text{ and}
τ\displaystyle\tau ={c1ct}\displaystyle=\{c_{1}\prec\dots\prec c_{t}\}

Where each ci=aij=bikc_{i}=a_{i_{j}}=b_{i_{k}} for some indices. Thus c1=aj=bkc_{1}=a_{j}=b_{k} for some indices, so guc1gu\prec c_{1} and vc1v\prec c_{1}, so {gu,v}c1\{gu,v\}\subset c_{1}. But c1c_{1} is a simplex of TT, so {gu,v}T\{gu,v\}\in T, as we claimed. ∎

7 Random 𝑮\bm{G}-Borsuk Graphs

In this section, we focus on the chromatic number of random GG-Borsuk graphs, and find thresholds for ε\varepsilon depending on nn, the number of vertices, such that asymptotically almost surely we can more precisely bound its chromatic number.

Definition 7.1 (Random GG-Borsuk graphs).

Let GG be a finite group, and KK a compact GG-space. We define the random GG-Graph on KK, denoted by 𝒢G(n,ε)\mathcal{G}_{G}\left(n,\varepsilon\right), as the GG-Borsuk graph 𝒢G(X(n),ε)\mathcal{G}_{G}\left(X(n),\varepsilon\right), where X(n)={x1,,xn}X(n)=\{x_{1},\dots,x_{n}\} is a set of nn i.i.d. uniform random points on KK.

Note that the definition of a random GG-Borsuk graph also depends on the specific underlying GG-space, however for simplicity we don’t include KK in the notation and instead make it clear from the context.

Since a random GG-Borsuk graph is an induced subgraph of a GG-Borsuk graph, Theorem 4.1 states its chromatic number must be at most covG(K)covG(k)\text{cov}_{G}\left(K\right)\leq\text{cov}_{G}\left(k\right) (here k=indG(K)k=\text{ind}_{G}\left(K\right)), with equality if the random sampling is dense enough. In this section we will prove Theorem 7.2, which establishes, up to a constant, what ε(n)\varepsilon(n) provides a dense enough sample. On the other hand, when ε(n)\varepsilon(n) is small enough to provide a sufficiently sparse sample, then the chromatic number is a.a.s. bounded above by covG(k1)\text{cov}_{G}\left(k-1\right), which if Conjecture 3.5 is true must be strictly less than covG(K)\text{cov}_{G}\left(K\right).

Theorem 7.2.

Consider a finite group GG and a finite geometric GG-simplicial complex KK of dimension dd. Suppose KK is pure dd-dimensional, i.e. all of its maximal faces have dimension dd. And let k=indG(K)k=\text{ind}_{G}\left(K\right). Then, there exist constants C1C_{1} and C2C_{2} (independent from nn), such that:

  1. 1.

    If εC1(lognn)1/d\varepsilon\geq C_{1}\left(\frac{\log n}{n}\right)^{1/d}, then a.a.s. χ(𝒢G(n,ε))=covG(K).\chi(\mathcal{G}_{G}\left(n,\varepsilon\right))=\text{cov}_{G}\left(K\right).

  2. 2.

    If εC2(lognn)1/k\varepsilon\leq C_{2}\left(\frac{\log n}{n}\right)^{1/k}, then a.a.s. χ(𝒢G(n,ε))covG(k1).\chi(\mathcal{G}_{G}\left(n,\varepsilon\right))\leq\text{cov}_{G}\left(k-1\right).

Theorem 3.9 follows by taking KK a EdGE_{d}G space, since then k=indG(K)=dim(K)=dk=\text{ind}_{G}\left(K\right)=\dim(K)=d.

The case G=2G=\mathbb{Z}_{2} corresponds to Random Borsuk Graphs, which was already established in Theorems 1.1 and 1.2 of [KMF20]. The main ideas behind the proofs of this generalization remain the same. We first establish some intermediate results before jumping into the proof.

7.1 Some Geometric Lemmas

For the “sparse case”, our approach relies on being able to project the vertices of the GG-Borsuk graph onto the (d1)(d-1)-skeleton of the classifying space via a GG-map. This will be possible as long as vertices that where close to each other remain close after the projection. This will follow from the following lemmas studying such projections on planes and dd-simplices.

Lemma 7.3.

Let OO be a point on the plane and let \ell be a line such that the distance from OO to \ell is at least d1>0d_{1}>0. Let AA and BB be two points such that the following hold

  • A,BA,B and OO lie all on the same semi-plane with respect to \ell;

  • the rays OA\vec{OA} and OB\vec{OB} intersect \ell;

  • 𝕕(O,A),𝕕(O,B)d2\mathbbm{d}\left(O,A^{\prime}\right),\mathbbm{d}\left(O,B^{\prime}\right)\leq d_{2}, where A=OAA^{\prime}=OA\cap\ell and B=OBB^{\prime}=OB\cap\ell;

  • 𝕕(O,A),𝕕(O,B)κε\mathbbm{d}\left(O,A\right),\mathbbm{d}\left(O,B\right)\geq\kappa\varepsilon; and

  • 𝕕(A,B)ε\mathbbm{d}\left(A,B\right)\leq\varepsilon.

Then

𝕕(A,B)d22κd1.\mathbbm{d}\left(A^{\prime},B^{\prime}\right)\leq\frac{d_{2}^{2}}{\kappa d_{1}}.
Proof.

Note

sin(OAB)=𝕕(O,)𝕕(O,A)d1d2.\sin(\measuredangle OA^{\prime}B^{\prime})=\frac{\mathbbm{d}\left(O,\ell\right)}{\mathbbm{d}\left(O,A^{\prime}\right)}\geq\frac{d_{1}}{d_{2}}.

By Law of Sines on the triangle ΔAOB\Delta AOB,

sin(AOB)=𝕕(A,B)sin(OBA)𝕕(O,A)𝕕(A,B)𝕕(O,A)εκε=1κ.\sin(\measuredangle AOB)=\frac{\mathbbm{d}\left(A,B\right)\sin(\measuredangle OBA)}{\mathbbm{d}\left(O,A\right)}\leq\frac{\mathbbm{d}\left(A,B\right)}{\mathbbm{d}\left(O,A\right)}\leq\frac{\varepsilon}{\kappa\varepsilon}=\frac{1}{\kappa}.

Note AOB=AOB\measuredangle A^{\prime}OB^{\prime}=\measuredangle AOB, so by Law of Sines on the triangle ΔAOB\Delta A^{\prime}OB^{\prime} we get

𝕕(A,B)=𝕕(O,B)sin(AOB)sin(OAB)d21κd1d2=d22κd1\mathbbm{d}\left(A^{\prime},B^{\prime}\right)=\frac{\mathbbm{d}\left(O,B^{\prime}\right)\sin(\measuredangle A^{\prime}OB^{\prime})}{\sin(\measuredangle OA^{\prime}B^{\prime})}\leq\frac{d_{2}\frac{1}{\kappa}}{\frac{d_{1}}{d_{2}}}=\frac{d_{2}^{2}}{\kappa d_{1}}

Lemma 7.4.

Let Δ\Delta be a geometric dd-simplex of diameter at most d2d_{2}. Let OO be an interior point such that 𝕕(Δ,O)d1\mathbbm{d}\left(\partial\Delta,O\right)\geq d_{1}. Take any two points A,BΔA,B\in\Delta with 𝕕(A,B)ε\mathbbm{d}\left(A,B\right)\leq\varepsilon and 𝕕(O,A),𝕕(O,B)κε\mathbbm{d}\left(O,A\right),\mathbbm{d}\left(O,B\right)\geq\kappa\varepsilon. Let AA^{\prime} and BB^{\prime} be the intersection of the rays OAOA and OBOB with the boundary of Δ\Delta respectively. Then 𝕕Δ(A,B)2(d+1)d22/(κd1)\mathbbm{d}_{\partial\Delta}\left(A^{\prime},B^{\prime}\right)\leq 2(d+1)d_{2}^{2}/(\kappa d_{1}).

Proof.

The case d=1d=1 is trivial, so we suppose d2d\geq 2. Let τ1\tau_{1} and τ2\tau_{2} be the (d1)(d-1)-faces of Δ\Delta that intersect the rays OA\vec{OA} and OB\vec{OB}, respectively. So Aτ1A^{\prime}\in\tau_{1} and Bτ2B^{\prime}\in\tau_{2}. We distinguish two cases:

  1. 1.

    If τ1=τ2\tau_{1}=\tau_{2}. Denote τ=τ1=τ2\tau=\tau_{1}=\tau_{2}, so both AA^{\prime} and BB^{\prime} lie on the same (d1)(d-1)-face τ\tau. In this case, let Π\Pi be the plane through O,AO,A and BB. The intersection τΠ\tau\cap\Pi must be a line segment containing AA^{\prime} and BB^{\prime}. Note that if \ell is the line containing τΠ\tau\cap\Pi, then O,A,BO,A,B and \ell satisfy all the hypothesis of Lemma 7.3, so

    𝕕Δ(A,B)d22κd12(d+1)d22κd1.\mathbbm{d}_{\partial\Delta}\left(A^{\prime},B^{\prime}\right)\leq\frac{d_{2}^{2}}{\kappa d_{1}}\leq 2(d+1)\frac{d_{2}^{2}}{\kappa d_{1}}.
  2. 2.

    If τ1τ2\tau_{1}\neq\tau_{2}. Consider the line segment AB¯\overline{AB}, which lies in the interior of Δ\Delta, and let p:AB¯Δp:\overline{AB}\to\partial\Delta be the projection on Δ\partial\Delta from OO, i.e. the intersection of rays OX\vec{OX} for all points XAB¯X\in\overline{AB}. Note p(AB¯)p(\overline{AB}) will be a polygonal line from Aτ1A^{\prime}\in\tau_{1} to Bτ2B^{\prime}\in\tau_{2}, passing through faces τ1=σ1,σ2,,σr=τ2\tau_{1}=\sigma_{1},\sigma_{2},\dots,\sigma_{r}=\tau_{2} of Δ\partial\Delta, where r(d+1)r\leq(d+1). Let Ei=p(AB¯)σiσi+1E^{\prime}_{i}=p(\overline{AB})\cap\sigma_{i}\cap\sigma_{i+1} for each i=1,,r1i=1,\dots,r-1, and say E0=AE_{0}^{\prime}=A^{\prime} and Er=BE_{r}^{\prime}=B^{\prime}. Correspondingly, let EiAB¯E_{i}\in\overline{AB} such that p(Ei)=Eip(E_{i})=E_{i}^{\prime}. For each i=1,,ri=1,\dots,r, let Πi\Pi_{i} be the plane through Ei1,EiE_{i-1},E_{i} and OO. As in the first case, let i\ell_{i} be the line containing Πiσi\Pi_{i}\cap\sigma_{i}. Clearly 𝕕(Ei1,Ei)ε\mathbbm{d}\left(E_{i-1},E_{i}\right)\leq\varepsilon, and, given κ>1\kappa>1, we also know 𝕕(AB¯,O)κε2\mathbbm{d}\left(\overline{AB},O\right)\geq\frac{\kappa\varepsilon}{2}, so 𝕕(Ei1,O),𝕕(Ei,O)κε/2\mathbbm{d}\left(E_{i-1},O\right),\mathbbm{d}\left(E_{i},O\right)\geq\kappa\varepsilon/2, then these points and i\ell_{i} satisfy the hypothesis of Lemma 7.3 with κ/2\kappa/2, so 𝕕Δ(Ei1,Ei)2d22κd1\mathbbm{d}_{\partial\Delta}\left(E_{i-1}^{\prime},E_{i}^{\prime}\right)\leq\frac{2d_{2}^{2}}{\kappa d_{1}}. Then

    𝕕Δ(A,B)𝕕Δ(E0,E1)++𝕕Δ(Er1,Er)r2d22κd12(d+1)d22κd1.\mathbbm{d}_{\partial\Delta}\left(A^{\prime},B^{\prime}\right)\leq\mathbbm{d}_{\partial\Delta}\left(E_{0}^{\prime},E_{1}^{\prime}\right)+\dots+\mathbbm{d}_{\partial\Delta}\left(E_{r-1}^{\prime},E_{r}^{\prime}\right)\leq r\frac{2d_{2}^{2}}{\kappa d_{1}}\leq 2\frac{(d+1)d_{2}^{2}}{\kappa d_{1}}.

We will use the next lemma to translate the volume of balls on EkGE_{k}G back to KK, where the random points are drawn.

Lemma 7.5.

Let σ\sigma be a geometric dd-dimensional simplex, τ\tau a kk-face of σ\sigma, and Ψ:στ\Psi:\sigma\to\tau a simplicial map that leaves τ\tau invariant. Then there exists a constant CC such that for each measurable subset EσE\subset\sigma, we have

Vold(E)CVolk(Ψ(E)).\text{Vol}_{d}\!\left(E\right)\leq C\text{Vol}_{k}\!\left(\Psi(E)\right).
Proof.

Let e1,e2,,ede_{1},e_{2},\dots,e_{d} be the standard vector basis of d\mathbb{R}^{d}, and let e0=0e_{0}=\vec{0} be the origin. For each mdm\leq d, the points {e0,,em}\{e_{0},\dots,e_{m}\} are affinely independent, so their convex hull is a geometric mm-simplex which we denote Δm\Delta_{m}. For each rmr\leq m, let πr:ΔmΔr\pi_{r}:\Delta_{m}\to\Delta_{r} be the projection map to the first rr coordinates. Note πr\pi_{r} can be seen as the simplicial map given by the vertex mapping

πr(ei)={ei if ire0 if i>r.\pi_{r}(e_{i})=\begin{cases}e_{i}&\text{ if }i\leq r\\ e_{0}&\text{ if }i>r\end{cases}.

We start by establishing the result when k=d1k=d-1. Enumerate the vertices of σ\sigma, V(σ)={a0,,ad}V(\sigma)=\{a_{0},\dots,a_{d}\}, such that V(τ)={a0,,ad1}V(\tau)=\{a_{0},\dots,a_{d-1}\}. Since the map Ψ\Psi is determined by its mapping of the vertices, Ψ(ai)=ai\Psi(a_{i})=a_{i} for all 01d10\leq 1\leq d-1 and Ψ(ad)\Psi(a_{d}) coincides with one of the other vertices, so shuffling the labels if necessary, Ψ(ad)=a0\Psi(a_{d})=a_{0}.

Then, there exists a unique affine transformation Φ:σΔd\Phi:\sigma\to\Delta_{d} such that Φ(ai)=ei\Phi(a_{i})=e_{i} for 0id0\leq i\leq d. Let φ=Φ|τ\varphi=\Phi|_{\tau}, note then φ:τΔd1\varphi:\tau\to\Delta_{d-1} is an affine map. Moreover, both Φ\Phi and φ\varphi are bijective, so there exist invertible matrices A,AA,A^{\prime} and vectors b,bb,b^{\prime} such that Φ(x)=Ax+b\Phi(x)=Ax+b and φ(x)=Ax+b\varphi(x)=A^{\prime}x+b^{\prime}. We then get that the following diagram of simplicial maps commutes

σ{\sigma}τ{\tau}Δd{\Delta_{d}}Δd1{\phantom{x}\Delta_{d-1}}Φ\scriptstyle{\Phi}Ψ\scriptstyle{\Psi}φ\scriptstyle{\varphi}φ1\scriptstyle{\varphi^{-1}}π\scriptstyle{\pi}\circlearrowleft

where π=πd1\pi=\pi_{d-1}. In particular, we have φ1πΦ=Ψ\varphi^{-1}\circ\pi\circ\Phi=\Psi. Let us also point out that for any set FΔdF\subset\Delta_{d}, we have Fπ(F)×[0,1]F\subset\pi(F)\times[0,1].

Then, the volume of any measurable set EσE\in\sigma satisfies:

Vold(E)\displaystyle\text{Vol}_{d}\!\left(E\right) =Vold(ΦE)|detA|\displaystyle=\frac{\text{Vol}_{d}\!\left(\Phi E\right)}{|\det A|}
Vold((πΦE)×[0,1])|detA|=Vold1(πΦE)1|detA|\displaystyle\leq\frac{\text{Vol}_{d}\!\left((\pi\Phi E)\times[0,1]\right)}{|\det A|}=\frac{\text{Vol}_{d-1}\!\left(\pi\Phi E\right)\cdot 1}{|\det A|}
=|detA||detA|Vold1(φ1πΦE)=CVold1(ΨE).\displaystyle=\frac{|\det A^{\prime}|}{|\det A|}\text{Vol}_{d-1}\!\left(\varphi^{-1}\pi\Phi E\right)=C\text{Vol}_{d-1}\!\left(\Psi E\right).

We now extend the result for any k<dk<d. Suppose V(σ)={a0,,ad}V(\sigma)=\{a_{0},\dots,a_{d}\} and V(τ)={a0,,ak}V(\tau)=\{a_{0},\dots,a_{k}\}. For each k+1idk+1\leq i\leq d, define the ii-face τiσ\tau_{i}\subset\sigma by τi={a0,,ak,ak+1,,ai}\tau_{i}=\{a_{0},\dots,a_{k},a_{k+1},\dots,a_{i}\}, so τk=τ\tau_{k}=\tau and τd=σ\tau_{d}=\sigma. Also for each k+1idk+1\leq i\leq d, define the map Ψi:τiτi1\Psi_{i}:\tau_{i}\to\tau_{i-1} as the simplicial map given by

Ψi(aj)={aj, if 0ji1Ψ(ai), if j=i.\Psi_{i}(a_{j})=\begin{cases}a_{j},\text{ if }0\leq j\leq i-1\\ \Psi(a_{i}),\text{ if }j=i\end{cases}.

Thus Ψ\Psi factors through the faces τi\tau_{i} as depicted in the following commutative diagram

σ=τd{\sigma=\tau_{d}}τd1{\tau_{d-1}}{\cdots}τk=τ{\tau_{k}=\tau}Ψd\scriptstyle{\Psi_{d}}Ψ\scriptstyle{\Psi}Ψd1\scriptstyle{\Psi_{d-1}}Ψk+1\scriptstyle{\Psi_{k+1}}

Finally, each Ψi:τiτi1\Psi_{i}:\tau_{i}\to\tau_{i-1} satisfies the hypothesis of the theorem for the special case that we already proved, so for each Eσ=τdE\subset\sigma=\tau_{d} measurable set, we get

Vold(E)\displaystyle\text{Vol}_{d}\!\left(E\right) CdVold1(Ψd(E))\displaystyle\leq C_{d}\text{Vol}_{d-1}\!\left(\Psi_{d}(E)\right)
CdCd1Vold2(Ψd1Ψd(E))\displaystyle\leq C_{d}C_{d-1}\text{Vol}_{d-2}\!\left(\Psi_{d-1}\Psi_{d}(E)\right)
\displaystyle\vdots
(i=k+1dCi)Volk(ΨdΨd1Ψk+1(E))=CVolk(Ψ(E))\displaystyle\leq\left(\prod_{i=k+1}^{d}C_{i}\right)\text{Vol}_{k}\!\left(\Psi_{d}\Psi_{d-1}\dots\Psi_{k+1}(E)\right)=C\text{Vol}_{k}\!\left(\Psi(E)\right)

7.2 Other Preliminary Results

We state the following Lemma without proof. While the proof should be straightforward, it is quite technical so we omit it. Refer to [BH99] for more on geodesics of geometric simplicial complexes, with the intrinsic metric.

Lemma 7.6.

Let KK be a geometric simplicial complex and ε0\varepsilon_{0} small. Then, there exists a constant NN, depending only on KK, such that for any x,yKx,y\in K with 𝕕(x,y)ε0\mathbbm{d}\left(x,y\right)\leq\varepsilon_{0}, there is a polygonal path from xx to yy of length 𝕕(x,y)\mathbbm{d}\left(x,y\right), made up of at most NN line segments, each of which is contained in a face of KK.

Theorem 7.7.

Let GG be a finite group, and KK a finite geometric GG-simplicial complex of dimension dd, such that GG acts on KK via isometries. Moreover, suppose KK is pure dd-dimensional, and let L=K(d1)L=K^{(d-1)} its (d1)(d-1)-skeleton. Fix a constant d1>0d_{1}>0, and let X0KX_{0}\subset K be a collection of one point in the interior of each dd-face of KK, such that gX0=X0gX_{0}=X_{0} for all gGg\in G, and 𝕕(X0,L)d1\mathbbm{d}\left(X_{0},L\right)\geq d_{1}. Given ε0>0\varepsilon_{0}>0, there exists a constant κ\kappa depending only on KK, ε0\varepsilon_{0} and d1d_{1}, such that for any ε\varepsilon small enough, there is a graph homomorphism

Φ:𝒢G(KB(X0,κε),ε)𝒢G(L,ε0).\Phi:\mathcal{G}_{G}\left(K\setminus B(X_{0},\kappa\varepsilon),\varepsilon\right)\to\mathcal{G}_{G}\left(L,\varepsilon_{0}\right).

Here B(X0,κε):=xX0B(x,κε)B(X_{0},\kappa\varepsilon):=\bigcup_{x\in X_{0}}B(x,\kappa\varepsilon).

Proof.

For each dd-face σiK\sigma_{i}\in K, let xiX0x_{i}\in X_{0} be the point in σi\sigma_{i}. For xσi{xi}x\in\sigma_{i}\setminus\{x_{i}\}, define φi(x)\varphi_{i}(x) to be the unique point of intersection of the ray xix\vec{x_{i}x} with the boundary σi\partial\sigma_{i}. Note this map φi:σi{xi}σi\varphi_{i}:\sigma_{i}\setminus\{x_{i}\}\to\partial\sigma_{i} is continuous and it restricts to the identity on σi\partial\sigma_{i}. Moreover, for any gGg\in G, since gg acts as a linear map, the ray xix\vec{x_{i}x} is mapped to the ray g(xi)g(x)\vec{g(x_{i})g(x)}, so if g(xi)=xjX0g(x_{i})=x_{j}\in X_{0}, we get

g(φi(x))=φj(g(x)).g\left(\varphi_{i}(x)\right)=\varphi_{j}\left(g(x)\right).

Let KdK_{d} be the collection of dd-faces of KK, and let d2=maxσKddiam(σ)d_{2}=\max_{\sigma\in K_{d}}\text{diam}\left(\sigma\right). Let NN be the constant from Lemma 7.6 corresponding to ε0\varepsilon_{0}. Define κ=2N(d+1)d22d1ε0\kappa=2\frac{N(d+1)d_{2}^{2}}{d_{1}\varepsilon_{0}}. Let ε<ε0\varepsilon<\varepsilon_{0} be small enough so that B(xi,κε)interior(σi)B(x_{i},\kappa\varepsilon)\subset\text{interior}(\sigma_{i}) for all ii. Then, by Lemma 7.4, for each ii and x,yσiB(xi,κε)x,y\in\sigma_{i}\setminus B(x_{i},\kappa\varepsilon) such that 𝕕K(x,y)ε\mathbbm{d}_{K}\left(x,y\right)\leq\varepsilon we get 𝕕L(φi(x),φi(y))ε0/N\mathbbm{d}_{L}\left(\varphi_{i}(x),\varphi_{i}(y)\right)\leq\varepsilon_{0}/N.

Define Φ:KB(X0,κε)L\Phi:K\setminus B(X_{0},\kappa\varepsilon)\to L by pasting all the maps φi\varphi_{i}. That is, if xint(σi)x\in\text{int}\left(\sigma_{i}\right) for some ii, let Φ(x)=φi(x)\Phi(x)=\varphi_{i}(x), and if xLx\in L, let Φ(x)=x\Phi(x)=x. Since all the maps φi\varphi_{i} are continuous and they all restrict to the identity on LL, pasting them produces a continuous map. Moreover, Φ\Phi is a GG-map:

Φ(g(x))=φi(g(x))=g(φj(x))=gΦ(x),\Phi(g(x))=\varphi_{i}(g(x))=g(\varphi_{j}(x))=g\Phi(x),

where ii and jj are the indices such that g(x)σig(x)\in\sigma_{i} and xσjx\in\sigma_{j}; and Φ(g(x))\Phi(g(x)) is defined since xσiB(xi,κε)x\in\sigma_{i}\setminus B(x_{i},\kappa\varepsilon) and gg is an isometry, so g(x)σjB(xj,κε)g(x)\in\sigma_{j}\setminus B(x_{j},\kappa\varepsilon).

We now consider the induced map

Φ:V(𝒢G(KB(X0,κε),ε))V(𝒢G(L,ε0)),\Phi:V(\mathcal{G}_{G}\left(K\setminus B(X_{0},\kappa\varepsilon),\varepsilon\right))\to V(\mathcal{G}_{G}\left(L,\varepsilon_{0}\right)),

by checking that it respects the adjacency of vertices, we will have a graph homomorphism as desired. Note that to do this is enough to prove that 𝕕L(Φ(x),Φ(y))ε0\mathbbm{d}_{L}\left(\Phi(x),\Phi(y)\right)\leq\varepsilon_{0} whenever 𝕕K(x,y)ε\mathbbm{d}_{K}\left(x,y\right)\leq\varepsilon, and x,yKB(X0,κε)x,y\in K\setminus B(X_{0},\kappa\varepsilon). Indeed, if xyx\sim y in 𝒢G(KB(x0,κε),ε)\mathcal{G}_{G}\left(K\setminus B(x_{0},\kappa\varepsilon),\varepsilon\right) then there is a gGg\in G, g𝟙g\neq\mathbbm{1}, such that 𝕕K(x,gy),xKB(X0,κε)\mathbbm{d}_{K}\left(x,gy\right),x\in K\setminus B(X_{0},\kappa\varepsilon) and since gg is an isometry, gyKB(X0,κε)gy\in K\setminus B(X_{0},\kappa\varepsilon). Thus if our claim holds, 𝕕L(Φ(x),gΦ(y))=𝕕L(Φ(x),Φ(gy))ε0\mathbbm{d}_{L}\left(\Phi(x),g\Phi(y)\right)=\mathbbm{d}_{L}\left(\Phi(x),\Phi(gy)\right)\leq\varepsilon_{0} and Φ(x)Φ(y)\Phi(x)\sim\Phi(y) in 𝒢G(L,ε0)\mathcal{G}_{G}\left(L,\varepsilon_{0}\right) as desired.

We now prove the claim. Let x,yKB(X0,κε)x,y\in K\setminus B(X_{0},\kappa\varepsilon), such that 𝕕K(x,y)ε<ε0\mathbbm{d}_{K}\left(x,y\right)\leq\varepsilon<\varepsilon_{0}. Then, there is a shortest polygonal path in KK between xx and yy of length less than ε\varepsilon. This path passes through the dd-faces τ1,τ2,,τr\tau_{1},\tau_{2},\dots,\tau_{r}, where xinterior(τ1)x\in\text{interior}(\tau_{1}), yinterior(τr)y\in\text{interior}(\tau_{r}) and rNr\leq N by Lemma 7.6. For each i=1,,r1i=1,\dots,r-1, let ziz_{i} be the intersection of this shortest path with τiτi+1\tau_{i}\cap\tau_{i+1}, let z0=xz_{0}=x and zr=yz_{r}=y. Note 𝕕K(zi,zi+1)ε\mathbbm{d}_{K}\left(z_{i},z_{i+1}\right)\leq\varepsilon for all i=0,r1i=0,\dots r-1, and xiLx_{i}\in L for i0,r+1i\neq 0,r+1, so 𝕕K(X0,zi)>κε\mathbbm{d}_{K}\left(X_{0},z_{i}\right)>\kappa\varepsilon.

Take i{1,,r}i\in\{1,\dots,r\}. If τi=σj\tau_{i}=\sigma_{j} is a dd-face of KK, as we noted before zi1,ziσjB(xj,κε)z_{i-1},z_{i}\in\sigma_{j}\setminus B(x_{j},\kappa\varepsilon) and 𝕕K(zi1,zi)ε\mathbbm{d}_{K}\left(z_{i-1},z_{i}\right)\leq\varepsilon, so 𝕕L(Φ(zi1),Φ(zi))=𝕕L(φ(zi1),φ(zi))2(d+1)d22κd1ε0N\mathbbm{d}_{L}\left(\Phi(z_{i-1}),\Phi(z_{i})\right)=\mathbbm{d}_{L}\left(\varphi(z_{i-1}),\varphi(z_{i})\right)\leq\frac{2(d+1)d_{2}^{2}}{\kappa d_{1}}\leq\frac{\varepsilon_{0}}{N}. If instead τiL\tau_{i}\in L, there is some dd-face σj\sigma_{j} such that τiσj\tau_{i}\leq\sigma_{j}, and again zi1,ziσjB(xj,κε)z_{i-1},z_{i}\in\sigma_{j}\setminus B(x_{j},\kappa\varepsilon) and𝕕K(zi1,zi)=ε\mathbbm{d}_{K}\left(z_{i-1},z_{i}\right)=\varepsilon, so 𝕕L(Φ(zi1),Φ(zi))ε0N\mathbbm{d}_{L}\left(\Phi(z_{i-1}),\Phi(z_{i})\right)\leq\frac{\varepsilon_{0}}{N}. Therefore,

𝕕L(Φ(x),Φ(y))i=1r𝕕L(Φ(zi1),Φ(zi))ε0,\mathbbm{d}_{L}\left(\Phi(x),\Phi(y)\right)\leq\sum_{i=1}^{r}\mathbbm{d}_{L}\left(\Phi(z_{i-1}),\Phi(z_{i})\right)\leq\varepsilon_{0},

which finishes the proof. ∎

Recall that a 𝜹\bm{\delta}-net \mathcal{B} of a compact metric space KK, is a finite subset such that for each xKx\in K, there is a yy\in\mathcal{B} such that 𝕕K(x,y)<δ\mathbbm{d}_{K}\left(x,y\right)<\delta. Similarly, a 𝜹\bm{\delta}-apart set \mathcal{B} of KK, is a finite subset such that for any x,yx,y\in\mathcal{B} with xyx\neq y, 𝕕K(x,y)>δ\mathbbm{d}_{K}\left(x,y\right)>\delta. The next lemma bounds the cardinality of δ\delta-nets and δ\delta-apart sets for geometric simplices, the technique mimics what we did for dd-Spheres on [KMF20].

Lemma 7.8.

Let Δ\Delta be a geometric dd-simplex. Then there exist constants C1(Δ)C_{1}(\Delta) and C2(Δ)C_{2}(\Delta) such that for every δ>0\delta>0 small enough, there is a δ\delta-net \mathcal{B} that is also a δ\delta-apart set, such that

C1(Δ)δd||C2(Δ)δd\frac{C_{1}(\Delta)}{\delta^{d}}\leq|\mathcal{B}|\leq\frac{C_{2}(\Delta)}{\delta^{d}}
Proof.

Regard Δ\Delta as embedded in d\mathbb{R}^{d}, with its incenter at the origin. Let δ>0\delta>0 small. Since Δ\Delta is compact, we can construct \mathcal{B} inductively. Indeed, choose any point y1Δy_{1}\in\Delta. Then, for each n2n\geq 2, if Bn=i=1nB(yi,δ)ΔB_{n}=\cup_{i=1}^{n}B(y_{i},\delta)\subsetneq\Delta, choose any yn+1ΔBmy_{n+1}\in\Delta B_{m}. Otherwise, stop and let ={y1,y2,,yn}\mathcal{B}=\{y_{1},y_{2},\dots,y_{n}\}. The process stops by compactness. Clearly \mathcal{B} is a δ\delta-net and a δ\delta-apart set. Note ΔB(yi,δ)\Delta\subset\bigcup B(y_{i},\delta), thus

Vol(Δ)Vol(B(yi,δ))\displaystyle\text{Vol}\!\left(\Delta\right)\leq\text{Vol}\!\left(\bigcup B(y_{i},\delta)\right) ||Vol(B(yi,δ))=||ωdδd, so\displaystyle\leq|\mathcal{B}|\text{Vol}\!\left(B(y_{i},\delta)\right)=|\mathcal{B}|\omega_{d}\delta^{d}\text{, so}
Vol(Δ)ωdδd\displaystyle\frac{\text{Vol}\!\left(\Delta\right)}{\omega_{d}\delta^{d}} ||.\displaystyle\leq|\mathcal{B}|.

Here ωd\omega_{d} is the volume of the unit dd-ball. If δ\delta is smaller than the inradius of Δ\Delta, we always have B(x,δ/2)2ΔB(x,\delta/2)\subset 2\Delta, for all xΔx\in\Delta. Since the yisy_{i}^{\prime}s are a δ\delta-apart set, clearly B(yi,δ/2)B(yj,δ/2)=B(y_{i},\delta/2)\cap B(y_{j},\delta/2)=\emptyset whenever iji\neq j, and so we have

2dVol(Δ)=Vol(2Δ)Vol(B(yi,δ/2))\displaystyle 2^{d}\text{Vol}\!\left(\Delta\right)=\text{Vol}\!\left(2\Delta\right)\geq\text{Vol}\!\left(\biguplus B(y_{i},\delta/2)\right) =||Vol(B(yi,δ/2))=||ωdδd2d, so\displaystyle=|\mathcal{B}|\text{Vol}\!\left(B(y_{i},\delta/2)\right)=|\mathcal{B}|\frac{\omega_{d}\delta^{d}}{2^{d}}\text{, so}
4dVol(Δ)ωdδd\displaystyle\frac{4^{d}\text{Vol}\!\left(\Delta\right)}{\omega_{d}\delta^{d}} ||.\displaystyle\geq|\mathcal{B}|.

Thus setting C1(Δ)=Vol(Δ)ωdC_{1}(\Delta)=\frac{\text{Vol}\!\left(\Delta\right)}{\omega_{d}} and C2(Δ)=4dVol(Δ)ωdC_{2}(\Delta)=\frac{4^{d}\text{Vol}\!\left(\Delta\right)}{\omega_{d}}, we get the result. ∎

Lemma 7.9.

Let GG be a finite group and KK a finite geometric GG-simplicial complex. Denote k=indG(K)dim(K)=dk=\text{ind}_{G}\left(K\right)\leq\text{dim}(K)=d. Then, there exists a Lipschitz GG-simplicial map Φ:K𝐺EkG\Phi:K\xrightarrow{G}E_{k}G and a constant C¯\overline{C} such that

Vold(Φ1(B(x,δ)))C¯Volk(B(x,δ))\text{Vol}_{d}\!\left(\Phi^{-1}\left(B(x,\delta)\right)\right)\leq\overline{C}\text{Vol}_{k}\!\left(B(x,\delta)\right)

for all xx and δ\delta such that B(x,δ)B(x,\delta) is contained in the interior of a kk-face of EkGE_{k}G.

Proof.

Since k=indG(K)k=\text{ind}_{G}\left(K\right), by definition there is some GG-map Φ:K𝐺EkG\Phi:K\xrightarrow{G}\ E_{k}G. Moreover, by the simplicial approximation theorem, we may assume that Φ\Phi is a simplicial map, substituting KK and EkGE_{k}G by some finer subdivisions if necessary. While the more general simplicial approximation theorem doesn’t guarantee that we still get a GG-map, this can be achieved by the standard procedure of defining it on a representative of each GG-orbit and then extending it appropriately (see e.g. [Mat08, Koz08]). Since a simplicial map induces a map between the geometric complexes via linear extensions, Φ:K𝐺EkG\Phi:K\xrightarrow{G}E_{k}G is just a piece-wise linear map, determined by the images of the vertices of KK. As such, and since linear maps are continuous iff they are bounded, Φ\Phi must be a Lipschitz map.

Suppose now that B(x,δ)B(x,\delta) is contained in the interior of kk-face τ\tau of EkGE_{k}G. Since Φ\Phi is a simplicial map, Φ1(τ)\Phi^{-1}(\tau) is a sub-simplicial complex of KK. Clearly the dd-volume of Φ1(B(x,δ))σ\Phi^{-1}(B(x,\delta))\cap\sigma is zero whenever dim(σ)<d\dim(\sigma)<d. At the same time, if σ\sigma is a face of Φ1(τ)\Phi^{-1}(\tau) such that dim(Φ(σ))<k\dim(\Phi(\sigma))<k, then B(x,δ)Φ(σ)=B(x,\delta)\cap\Phi(\sigma)=\emptyset so Φ1(B(x,δ))σ=\Phi^{-1}(B(x,\delta))\cap\sigma=\emptyset. Hence, the dd-volume of Φ1(B(x,δ))\Phi^{-1}(B(x,\delta)) only depends on it intersection with the dd-faces of Φ1(τ)\Phi^{-1}(\tau) that are mapped onto τ\tau surjectively.

Denote Kd(τ)={σK:dim(σ)=d and Φ(σ)=τ}K_{d}(\tau)=\left\{\sigma\in K:\dim(\sigma)=d\text{ and }\Phi(\sigma)=\tau\right\} and let σKd(τ)\sigma\in K_{d}(\tau). Enumerate σ={a0,a1,,ad}\sigma=\{a_{0},a_{1},\dots,a_{d}\} such that if ω={a0,,ak}\omega=\{a_{0},\dots,a_{k}\}, then Φ(ω)=τ\Phi(\omega)=\tau. Thus, for each aia_{i} with i>ki>k, Φ(ai)=Φ(aj)\Phi(a_{i})=\Phi(a_{j}) for exactly one index 0jk0\leq j\leq k. Let Ψ:σω\Psi:\sigma\to\omega be the simplicial map given by the assignment of vertices

Ψ(ai)={ai, if 0ikaj, s.t. Φ(ai)=Φ(aj), if k+1id,\Psi(a_{i})=\begin{cases}a_{i},\text{ if }0\leq i\leq k\\ a_{j},\text{ s.t. }\Phi(a_{i})=\Phi(a_{j}),\text{ if }k+1\leq i\leq d\end{cases},

and let φ=Φ|ω\varphi=\Phi|_{\omega} the restriction of Φ\Phi to ω\omega. Note that these two maps give the factorization Φ=Ψφ\Phi=\Psi\circ\varphi. Note also that φ\varphi is a bijective affine transformation, so there exists an invertible matrix AA and a vector bb such that φ(x)=Ax+b\varphi(x)=Ax+b. Let E(σ)=Φ1(B(x,δ))σE(\sigma)=\Phi^{-1}(B(x,\delta))\cap\sigma, clearly E(σ)E(\sigma) is a measurable subset, and applying Lemma 7.5 to σ\sigma, ω\omega and the map Ψ\Psi, there exists a constant C(σ)C(\sigma) such that

Vold(E)\displaystyle\text{Vol}_{d}\!\left(E\right) C(σ)Volk(Ψ(E))=C(σ)|detA|Volk(φΨ(E))\displaystyle\leq C(\sigma)\text{Vol}_{k}\!\left(\Psi(E)\right)=\frac{C(\sigma)}{|\det A|}\text{Vol}_{k}\!\left(\varphi\Psi(E)\right)
=C(σ)|detA|Volk(Φ(E))C(σ)|detA|Volk(B(x,δ))\displaystyle=\frac{C(\sigma)}{|\det A|}\text{Vol}_{k}\!\left(\Phi(E)\right)\leq\frac{C(\sigma)}{|\det A|}\text{Vol}_{k}\!\left(B(x,\delta)\right)
=C¯(σ)Volk(B(x,δ)).\displaystyle=\overline{C}(\sigma)\text{Vol}_{k}\!\left(B(x,\delta)\right).

Thus, letting

C^(τ)=σKd(τ)C¯(σ),andC¯=maxτEkGdim(τ)=kC^(τ),\hat{C}(\tau)=\sum_{\sigma\in K_{d}(\tau)}\overline{C}(\sigma),\phantom{xx}\text{and}\phantom{xx}\overline{C}=\max_{\begin{subarray}{c}\tau\in E_{k}G\\ \dim(\tau)=k\end{subarray}}\hat{C}(\tau),

we get the desired inequality. ∎

7.3 Proof of Theorem 7.2

7.3.1 Dense Case

If ε0\varepsilon\to 0 slowly enough, with high probability the random vertices will form a small net, and so Theorem 4.2 will guarantee that the chromatic number is the same as covG(K)\text{cov}_{G}\left(K\right). The argument behind is a union bound, and the independence of the random vertices.

Proof of dense case.

Note that as soon as ε(n)\varepsilon(n) is small enough, Theorem 4.2 item 2 gives the upper bound

χ(𝒢G(n,ε))covG(K).\chi(\mathcal{G}_{G}\left(n,\varepsilon\right))\leq\text{cov}_{G}\left(K\right).

Let Kd={σK|dim(σ)=d}K_{d}=\{\sigma\in K|\dim(\sigma)=d\} be the set of dd-faces of KK. Thus, since KK is pure dd-dimensional, we have K=σKdσ\lVert K\rVert=\left\lVert\bigcup_{\sigma\in K_{d}}\sigma\right\rVert. By Lemma 7.8, for each dd-face σKd\sigma\in K_{d}, there exist constants C¯(σ)\overline{C}(\sigma) such that for any given δ>0\delta>0 small, we can find a δ\delta-net (σ)\mathcal{B}(\sigma) of σ\sigma such that |(σ)|C¯(σ)δd|\mathcal{B}(\sigma)|\leq\overline{C}(\sigma)\delta^{-d}. Define a global C¯:=σKdC¯(σ)\overline{C}:=\sum_{\sigma\in K_{d}}\overline{C}(\sigma). Hence, by defining =σKd(σ)\mathcal{B}=\bigcup_{\sigma\in K_{d}}\mathcal{B}(\sigma) we can always find a δ\delta-net of KK such that ||C¯δd|\mathcal{B}|\leq\overline{C}\delta^{-d}.

Similarly, the volume of a dd-ball of radius δ\delta is proportional to δd\delta^{d}. And it is known that for a dd-face σ\sigma, if δ\delta is small enough and xσx\in\sigma then

V(B(x,δ)σ)c(σ)δd,V(B(x,\delta)\cap\sigma)\geq c(\sigma)\delta^{d},

for a constant c(σ)>0c(\sigma)>0 that depends on the hyper-angle of the sharpest corner of σ\sigma. Define c:=minσKdc(σ)V(K)c:=\frac{\min_{\sigma\in K_{d}}c(\sigma)}{V(K)}, thus for any point xKx\in K and δ>0\delta>0 small, we have

V(B(x,δ)K)V(K)V(B(x,δ)σ)V(K)c(σ)V(K)δdcδd.\frac{V(B(x,\delta)\cap K)}{V(K)}\geq\frac{V(B(x,\delta)\cap\sigma)}{V(K)}\geq\frac{c(\sigma)}{V(K)}\delta^{d}\geq c\delta^{d}.

Let DD be as in Theorem 4.2 item 1. Then, we claim that C1=2DcdC_{1}=\frac{2D}{\sqrt[d]{c}} is as stated in the theorem. That is, if εC1(logn/n)1/d\varepsilon\geq C_{1}(\log n/n)^{1/d}, then we will show that a.a.s. covG(K)χ(𝒢G(n,ε))\text{cov}_{G}\left(K\right)\leq\chi\left(\mathcal{G}_{G}\left(n,\varepsilon\right)\right).

For each nn, let δ=ε(n)/(2D)\delta=\varepsilon(n)/(2D), and consider the corresponding δ\delta-net of KK, ={y1,,yN}\mathcal{B}=\{y_{1},\dots,y_{N}\}. For each i=1,,Ni=1,\dots,N define the closed set Fi=B(yi,δ)KF_{i}=B(y_{i},\delta)\cap K. Then K=i=1NFiK=\cup_{i=1}^{N}F_{i}, V(Fi)V(K)cδd\frac{V(F_{i})}{V(K)}\geq c\delta^{d}, and NC¯δdC¯c(n/logn)N\leq\overline{C}\delta^{-d}\leq\overline{C}c(n/\log n) as discussed above.

Let X={x1,,xn}X=\{x_{1},\cdots,x_{n}\} be the vertices of the random GG-Borsuk graph, i.e. they are i.i.d. uniform random points on KK. The computations below will establish that a.a.s. XX is a (ε/D)net(\varepsilon/D)-net of KK, and thus Theorem 4.2 item 1 gives the desired lower bound for the chromatic number.

For any i=1,,Ni=1,\dots,N and j=1,,nj=1,\dots,n we have

[xjFi]\displaystyle\mathbb{P}\left[x_{j}\in F_{i}\right] =V(Fi)V(K)cδd=cεd(2D)dlognn.\displaystyle=\frac{V(F_{i})}{V(K)}\geq c\delta^{d}=c\frac{\varepsilon^{d}}{(2D)^{d}}\geq\frac{\log n}{n}.

Then, the probability that some of the sets FiF_{i} doesn’t contain any vertex of XX is

[i=1N(XFi=)]\displaystyle\mathbb{P}\left[\bigvee_{i=1}^{N}(X\cap F_{i}=\emptyset)\right] i=1N[XFi=]=i=1N[xjFi]n\displaystyle\leq\sum_{i=1}^{N}\mathbb{P}\left[X\cap F_{i}=\emptyset\right]=\sum_{i=1}^{N}\mathbb{P}\left[x_{j}\not\in F_{i}\right]^{n}
i=1N(1logn/n)nN(1logn/n)n\displaystyle\leq\sum_{i=1}^{N}(1-\log n/n)^{n}\leq N(1-\log n/n)^{n}
C¯cnlognelogn=C¯clogn0 as n.\displaystyle\leq\overline{C}c\frac{n}{\log n}e^{-\log n}=\frac{\overline{C}c}{\log n}\to 0\text{ as }n\to\infty.

Thus, a.a.s. every set FiF_{i} will contain some vertex of XX, and hence the set XX is a 2δ=(ε/D)2\delta=(\varepsilon/D)-net of KK as desired. ∎

7.3.2 Sparse Case

When ε0\varepsilon\to 0 fast, and if K=EkGK=E_{k}G, a.a.s. there will be a point x0Kx_{0}\in K, not too close to K(k1)K^{(k-1)}, such that its orbit under GG is always at a distance at least κε\kappa\varepsilon of all the random vertices of 𝒢G(n,ε)\mathcal{G}_{G}\left(n,\varepsilon\right). In this case, Theorem 7.7 provides a graph homomorphism from the random GG-Borsuk graph on KK to a GG-Borsuk graph on K(k1)K^{(k-1)}, and so Theorem 4.2 gives the desired upper bound. When KK is not a finite classifying space, the same will still be true by looking at the pre-images of the GG-map Φ:KEkG\Phi:K\to E_{k}G.

The idea behind the proof is to consider the pre-images of open balls around many orbits Gx0Gx_{0} in EkGE_{k}G, and prove that a.a.s. there will be at least one of them that doesn’t contain any vertex of 𝒢G(n,ε)\mathcal{G}_{G}\left(n,\varepsilon\right). To get spacial independence of random events, we consider a Poissonization of the random vertices, i.e. we draw them via a Poisson Point Process instead. We include a couple of results about Poisson Point Processes and the Poisson distribution without proof, for their proofs and a complete discussion refer to [Pen03] or [Kin93].

Theorem 7.10 (Poissonization).

Let x1,x2,x_{1},x_{2},\dots, be uniform random variables on a compact metric space KK. Let mPois(λ)m\sim\text{Pois}(\lambda) and let η\eta be the random counting measure associated to the point process PλP_{\lambda}= {x1,x2,,xm}\{x_{1},x_{2},\dots,x_{m}\}. Then PλP_{\lambda} is a Poisson Point Process and for a Borel AKA\subset K, η(A)Pois(λV(A)V(K))\eta(A)\sim\text{Pois}\left(\lambda\frac{V(A)}{V(K)}\right).

Lemma 7.11.

For n0n\geq 0, [Pois(2n)<n]e0.306n.\mathbb{P}\left[\text{Pois}\left(2n\right)<n\right]\leq e^{-0.306n}.

To make our proofs cleaner, we address the probabilistic part in the following lemma.

Lemma 7.12.

Let KK be a compact metric space pure dd-dimensional. Let DD be a constant such that for each nn, there are N=N(n)N=N(n) disjoint closed sets F1,,FNKF_{1},\dots,F_{N}\subset K such that

  1. 1.

    Vol(Fr)Vol(K)logn3n\frac{\text{Vol}\!\left(F_{r}\right)}{\text{Vol}\!\left(K\right)}\leq\frac{\log n}{3n} for all r=1,,Nr=1,\dots,N; and

  2. 2.

    NDnlognN\geq\frac{Dn}{\log n}.

Let Xn={x1,,xn}X_{n}=\{x_{1},\dots,x_{n}\} be a collection of nn i.i.d. uniform random points on KK with corresponding counting measure η1n\eta_{1}^{n}. Then

[r=1Nη1n(Fi)=0]1 as n.\mathbb{P}\left[\bigvee_{r=1}^{N}\eta_{1}^{n}(F_{i})=0\right]\to 1\text{ as }n\to\infty.
Proof.

We use the fact that the FrF_{r}’s are disjoint, by considering instead the Poissonization of the random points, so that we get spatial independence. Hence let x1,x2,x_{1},x_{2},\dots be uniform random variables on KK, let MPois(2n)M\sim\text{Pois}\left(2n\right), and consider the Poisson Point Process {x1,,xM}\{x_{1},\dots,x_{M}\}, as described in Theorem 7.10. Letting η\eta be the corresponding counting measure, we get

[r=1Nη(Fr)>0]\displaystyle\mathbb{P}\left[\bigwedge_{r=1}^{N}\eta(F_{r})>0\right] =r=1N[η(Fr)>0]\displaystyle=\prod_{r=1}^{N}\mathbb{P}\left[\eta(F_{r})>0\right]
=r=1N(1[Pois(2nVol(Fr)Vol(K))=0])\displaystyle=\prod_{r=1}^{N}\left(1-\mathbb{P}\left[\text{Pois}\left(\frac{2n\text{Vol}\!\left(F_{r}\right)}{\text{Vol}\!\left(K\right)}\right)=0\right]\right)
=r=1N(1e2nVol(Fr)Vol(K))(1e(2/3)logn)N\displaystyle=\prod_{r=1}^{N}\left(1-e^{-\frac{2n\text{Vol}\!\left(F_{r}\right)}{\text{Vol}\!\left(K\right)}}\right)\leq\left(1-e^{-(2/3)\log n}\right)^{N}
exp(Nn2/3)exp(Dn1/3logn)0 as n.\displaystyle\leq\exp\left(-Nn^{-2/3}\right)\leq\exp\left(-\frac{Dn^{1/3}}{\log n}\right)\to 0\text{ as }n\to\infty.

Thus [r=1Nη(Fr)=0]1\mathbb{P}\left[\bigvee_{r=1}^{N}\eta(F_{r})=0\right]\to 1 as nn\to\infty. Clearly when MnM\geq n, η(Fr)=0\eta(F_{r})=0 implies η1n(Fr)=0\eta_{1}^{n}(F_{r})=0 as well. Thus, applying Lemma 7.11, we get

[r=1Nη(Fr)=0]\displaystyle\mathbb{P}\left[\bigvee_{r=1}^{N}\eta(F_{r})=0\right] [r=1Nη1n(Fr)=0]+[M<n]\displaystyle\leq\mathbb{P}\left[\bigvee_{r=1}^{N}\eta_{1}^{n}(F_{r})=0\right]+\mathbb{P}\left[M<n\right]
[r=1Nη1n(Fr)=0]+o(1).\displaystyle\leq\mathbb{P}\left[\bigvee_{r=1}^{N}\eta_{1}^{n}(F_{r})=0\right]+o(1).

From where the desired result follows. ∎

Proof of Sparse Case.

Let MM be a geometric realization of EkGE_{k}G such that GG acts on MM via isometries. Then, let Φ:K𝐺M\Phi:K\xrightarrow{G}M be the simplicial GG-map and CC the corresponding constant given by Lemma 7.9. Let λ\lambda be the Lipschitz constant of Φ\Phi.

Let L=M(k1)L=M^{(k-1)} be the (k1)(k-1)-skeleton of MM. Restricting the action of GG to LL we also get a geometric GG-simplicial complex. Moreover, dim(L)=k1\dim(L)=k-1 and, since MM is kk-connected, LL must be (k1)(k-1)-connected, thus LL is a (k1)(k-1)-classifying space of GG, and covG(L)=covG(k1)\text{cov}_{G}\left(L\right)=\text{cov}_{G}\left(k-1\right). Let ε0>0\varepsilon_{0}>0 be as in Theorem 4.2, that is, for any XLX\subset L and any 0<εε00<\varepsilon\leq\varepsilon_{0},

χ(𝒢G(X,ε))covG(L)=covG(k1).\chi\left(\mathcal{G}_{G}\left(X,\varepsilon\right)\right)\leq\text{cov}_{G}\left(L\right)=\text{cov}_{G}\left(k-1\right).

Let MdM_{d} be the set of dd-faces of MM. Let d1=(1ρ)minσMdinradius(σ)d_{1}=(1-\rho)\min_{\sigma\in M_{d}}{\text{inradius}(\sigma)}, for ρ<1\rho<1 any fixed constant close to 1. And let κ\kappa be the constant given by Theorem 7.7, corresponding to ε0\varepsilon_{0} and d1d_{1}. Partition the dd-faces of MM into GG-orbits, and let 𝒞={σ1,σm}\mathcal{C}=\{\sigma_{1},\dots\sigma_{m}\} be a collection of one representative per orbit, where m=|𝒞|m=|\mathcal{C}|.

We will prove that the constant

C2=1κλ(Vold(K)3m|G|Cωk)1/k,C_{2}=\frac{1}{\kappa\lambda}\left(\frac{\text{Vol}_{d}\!\left(K\right)}{3m|G|C\omega_{k}}\right)^{1/k},

where ωk\omega_{k} is the volume of the unit kk-ball, is as stated in the theorem. That is, if ε(n)C2(logn/n)1/k\varepsilon(n)\leq C_{2}(\log n/n)^{1/k}, then a.a.s. χ(𝒢G(n,ε))covG(L)\chi(\mathcal{G}_{G}\left(n,\varepsilon\right))\leq\text{cov}_{G}\left(L\right).

For a dd-face σ\sigma, denote by ρσ\rho\sigma the homothetic copy of σ\sigma obtained by scaling it by a factor of ρ\rho with respect to its incenter. Note that

𝕕(ρσ,σ)=(1ρ)inradius(σ)d1.\mathbbm{d}\left(\rho\sigma,\partial\sigma\right)=(1-\rho)\text{inradius}(\sigma)\geq d_{1}.

Fix nn large enough so that ε\varepsilon satisfies κλεd1\kappa\lambda\varepsilon\ll d_{1}, and let δ=2κλε\delta=2\kappa\lambda\varepsilon. For each 1im1\leq i\leq m, Lemma 7.8 guarantees there is a constant C¯(ρσi)\overline{C}(\rho\sigma_{i}) and a δ\delta-apart set i={y1i,,y|i|i}ρσi\mathcal{B}_{i}=\{y_{1}^{i},\dots,y_{|\mathcal{B}_{i}|}^{i}\}\subset\rho\sigma_{i} such that,

C¯(ρσi)δk|i|.\frac{\overline{C}(\rho\sigma_{i})}{\delta^{k}}\leq|\mathcal{B}_{i}|.

Let C¯=miniC¯(ρσi)\overline{C}=\min_{i}{\overline{C}(\rho\sigma_{i})} and N=mini|i|N=\min_{i}|\mathcal{B}_{i}|, then C¯/δkN\overline{C}/\delta^{k}\leq N.

For each r=1,,Nr=1,\dots,N construct the set YrY_{r}, as

Yr:={gyri:gG,1im}.Y_{r}:=\{gy_{r}^{i}:g\in G,1\leq i\leq m\}.

Note this construction guarantees Yr=gYrY_{r}=gY_{r} for all gGg\in G, it has one point in the interior of each dd-face of MM, and 𝕕(Yr,L)d1\mathbbm{d}\left(Y_{r},L\right)\geq d_{1}. Thus, for each 1rN1\leq r\leq N, Theorem 7.7 guarantees that there is a graph homomorphism

𝒢G(MB(Yr,κλε),λε)𝒢G(L,ε0).\mathcal{G}_{G}\left(M\setminus B(Y_{r},\kappa\lambda\varepsilon),\lambda\varepsilon\right)\to\mathcal{G}_{G}\left(L,\varepsilon_{0}\right).

Note that because the points in YrY_{r} come from δ\delta-apart sets and κλε<d1\kappa\lambda\varepsilon<d_{1}, then B(Yr,κλε)B(Yr,κλε)=B(Y_{r},\kappa\lambda\varepsilon)\cap B(Y_{r^{\prime}},\kappa\lambda\varepsilon)=\emptyset whenever rrr\neq r^{\prime}, and B(Yr,κλε)B(Y_{r},\kappa\lambda\varepsilon) is a disjoint union of kk-balls.

For each r=1,Nr=1,\dots N, let Fr=Φ1(B(Yr,κλε))F_{r}=\Phi^{-1}\left(B(Y_{r},\kappa\lambda\varepsilon)\right), thus by Lemma 7.9, we get

Vold(Fr)\displaystyle\text{Vol}_{d}\!\left(F_{r}\right) =yYrVold(Φ1(B(y,κλε)))yYrCVolk(B(y,κλε))\displaystyle=\sum_{y\in Y_{r}}\text{Vol}_{d}\!\left(\Phi^{-1}\left(B(y,\kappa\lambda\varepsilon)\right)\right)\leq\sum_{y\in Y_{r}}C\text{Vol}_{k}\!\left(B(y,\kappa\lambda\varepsilon)\right)
=|Yr|Cωk(κλε)km|G|Cωk(κλC2)k(lognn)\displaystyle=|Y_{r}|C\omega_{k}(\kappa\lambda\varepsilon)^{k}\leq m|G|C\omega_{k}(\kappa\lambda C_{2})^{k}\left(\frac{\log n}{n}\right)
Vold(K)logn3n.\displaystyle\leq\frac{\text{Vol}_{d}\!\left(K\right)\log n}{3n}.

So Vold(Fr)Vold(K)logn3n\frac{\text{Vol}_{d}\!\left(F_{r}\right)}{\text{Vol}_{d}\!\left(K\right)}\leq\frac{\log n}{3n} for all r=1,Nr=1,\dots N, and we also have

NC¯(2κλε)kC¯(2κλC2)knlogn=Cnlogn.N\geq\frac{\overline{C}}{(2\kappa\lambda\varepsilon)^{k}}\geq\frac{\overline{C}}{(2\kappa\lambda C_{2})^{k}}\cdot\frac{n}{\log n}=C^{\prime}\frac{n}{\log n}.

Thus we can apply Lemma 7.12 to XnKX_{n}\subset K, nn i.i.d. uniform random points on KK, so a.a.s. there exists some index rr such that XnFr=X_{n}\cap F_{r}=\emptyset. Finally, since Φ\Phi is a Lipschitz GG-map, by Theorem 4.1 item 2, it induces a graph homomorphism, and so we have the following composition of graph homomorphisms

𝒢G(Xn,ε)𝒢G(KFr,ε)Φ𝒢G(MB(Yr,κλε),λε)𝒢G(L,ε0)\displaystyle\mathcal{G}_{G}\left(X_{n},\varepsilon\right)\hookrightarrow\mathcal{G}_{G}\left(K\setminus F_{r},\varepsilon\right)\xrightarrow{\Phi}\mathcal{G}_{G}\left(M\setminus B(Y_{r},\kappa\lambda\varepsilon),\lambda\varepsilon\right)\to\mathcal{G}_{G}\left(L,\varepsilon_{0}\right)

So we get the desired result

χ(𝒢G(n,ε))χ(𝒢G(L,ε0))covG(L)=covG(k1).\chi\left(\mathcal{G}_{G}\left(n,\varepsilon\right)\right)\leq\chi\left(\mathcal{G}_{G}\left(L,\varepsilon_{0}\right)\right)\leq\text{cov}_{G}\left(L\right)=\text{cov}_{G}\left(k-1\right).

8 Further Questions

  • While we have some computer-aided examples suggesting that Conjecture 3.5 might be true, it is still open. Thus, natural follow up questions would be to prove or disprove Conjectures 3.6 and 3.5. Towards this end, it would be interesting to produce GG-covers for more small groups in dimensions 2 and 3. In particular, it would be useful to be able to compute covm(2)\text{cov}_{\mathbb{Z}_{m}}\left(2\right) and covm(3)\text{cov}_{\mathbb{Z}_{m}}\left(3\right) with more efficient algorithms, rather than solving an integer programming problem.

  • For the random GG-Borsuk graphs, we wonder whether there exist similar thresholds to that in Theorem 3.9 for each possible chromatic number. We state this question as a conjecture.

    Conjecture 8.1.

    Let GG be a finite group, and KK a geometric EdGE_{d}G space. Then, for each 1kd+11\leq k\leq d+1, there exist constants Ck,C¯kC_{k},\overline{C}_{k} and functions fk:+f_{k}:\mathbb{N}\to\mathbb{R}^{+}, such that:

    If Ckfk(n)ε(n)C¯kfk+1(n)C_{k}f_{k}(n)\leq\varepsilon(n)\leq\overline{C}_{k}f_{k+1}(n), then a.a.s. χ(𝒢G(n,ε))=covG(k)\chi\left(\mathcal{G}_{G}\left(n,\varepsilon\right)\right)=\text{cov}_{G}\left(k\right).

    In particular, Theorem 3.9 shows the conjecture is true for k=dk=d, taking fd(n)=(logn/n)1/df_{d}(n)=\left(\log n/n\right)^{1/d} and fd+1(n)=ε0f_{d+1}(n)=\varepsilon_{0} from Theorem 4.2.

9 Acknowledgments

The author wants to thank his advisor Matthew Kahle, for fruitful conversations, frequent words of encouragement, and for kindly proof-reading and suggesting improvements to parts of this paper.

Appendix A Computer-aided Examples

In this appendix we describe the approach taken to compute the values of Table 1 and the graphs on Figure 1(d). Let us start by describing a general method for any finite group GG, and then we will discuss our specific approach in dimensions 2 and 3 for cyclic groups.

General approach

  1. 1.

    Suppose LL is a EdGE_{d}G classifying space where GG acts via isometries. Moreover, suppose we know covG(d)=|G|+d\text{cov}_{G}\left(d\right)=|G|+d and we have a corresponding closed GG-cover of LL.

  2. 2.

    Take K=GLK=G*L, so KK is a Ed+1GE_{d+1}G classifying space, and GG acts on KK via isometries.

  3. 3.

    Let Δ\Delta be a subdivision operation that respects the GG-action, that is, an algorithm that takes KK and produces a GG-simplicial complex ΔK\Delta K with the same geometric realization ΔK=K\|\Delta K\|=\|K\|, with V(K)V(ΔK)V(K)\subset V(\Delta K), and such that the action of GG restricted to ΔK\Delta K is again a GG-action via simplicial maps. Denote by ΔkK\Delta_{k}K the repeated application of Δ\Delta. Note that we can realize LKL\hookrightarrow K, so Δ\Delta restricted to LL also produces a subdivision of LL

  4. 4.

    For T=ΔkKT=\Delta_{k}K, let H(T)H(T) be the graph with vertices V(T)V(T) and edges uvu\sim v whenever there is a gGg\in G, g𝟙g\neq\mathbbm{1} s.t. {u,gv}T\{u,gv\}\in T. Similarly, we define H(T0)H(T_{0}) for T0=ΔkLT_{0}=\Delta_{k}L.

  5. 5.

    Let kk be the minimum integer such that the known GG-cover for LL produces a proper graph coloring for H(T0)H(T_{0}), where T0=ΔkLT_{0}=\Delta_{k}L. Thus χ(H(T0))=covG(d)=|G|+d\chi(H(T_{0}))=\text{cov}_{G}\left(d\right)=|G|+d.

  6. 6.

    Let c0:V(T0){1,,|G|+d}c_{0}:V(T_{0})\to\{1,\dots,|G|+d\} be this coloring.

  7. 7.

    Let T=ΔkKT=\Delta_{k}K, and H=H(T)H=H(T). We define a partial coloring c:V(T){1,|G|+d,|G|+d+1}c:V(T)\to\{1,\dots|G|+d,|G|+d+1\}, as follows

    c(v)={c0(v0) if v0TL|G|+d+1 if v=t𝟙(1t)v0 for some t>0,v0V(T0)c(v)=\begin{cases}c_{0}(v_{0})&\text{ if }v_{0}\in T\cap L\\ |G|+d+1&\text{ if }v=t\mathbbm{1}\oplus(1-t)v_{0}\text{ for some }t>0,v_{0}\in V(T_{0})\end{cases}
  8. 8.

    We then attempt to extend this partial coloring to the remaining vertices of HH: {tg(1t)v0:gG,g𝟙,v0V(T0),0<t1}\{tg\oplus(1-t)v_{0}:g\in G,g\neq\mathbbm{1},v_{0}\in V(T_{0}),0<t\leq 1\}. We do this by solving the corresponding integer programming problem. If this is possible, we continue with the next step. Otherwise, we increase kk by one and repeat steps 6-7.

  9. 9.

    If the partial coloring can be extended to a proper coloring for HH, Theorem 6.5 guarantees covG(d+1)χ(H)=|G|+d+1\text{cov}_{G}\left(d+1\right)\leq\chi(H)=|G|+d+1, and as in its proof, coloring the Barycentric Subdivision with the corresponding colors of the vertices of TT, provides a GG-cover for KK.

  10. 10.

    This whole process may be repeated for dimension d+2d+2, using the GG-cover of KK found as the starting point.

For Cyclic Groups

2D Case

Let m={𝟙,ν,ν2,,νm1}\mathbb{Z}_{m}=\{\mathbbm{1},\nu,\nu^{2},\dots,\nu^{m-1}\} be a cyclic group of order mm. Let CmC_{m} be the cycle of length mm, with its vertices labeled by the elements of m\mathbb{Z}_{m}. It is convenient to think of the rational points of CmC_{m} as labeled by elements in the module [m]\mathbb{Q}[\mathbb{Z}_{m}]. That is, the rational points in the segment [νi,νi+1][\nu^{i},\nu^{i+1}] can be described as ανi+(1α)νi+1\alpha\nu^{i}+(1-\alpha)\nu^{i+1}, for 0α10\leq\alpha\leq 1, α\alpha\in\mathbb{Q}. Thus, taking left multiplication in the module [m]\mathbb{Q}[\mathbb{Z}_{m}] corresponds to the m\mathbb{Z}_{m}-action on Cm\|C_{m}\|.

  1. 1.

    We take E1m=𝕊1=CmE_{1}\mathbb{Z}_{m}=\mathbb{S}^{1}=C_{m}, so covm(1)=m+1\text{cov}_{\mathbb{Z}_{m}}\left(1\right)=m+1 and can be realized by dividing 𝕊1\mathbb{S}^{1} into (m+1)(m+1)-arcs of the same size.

  2. 2.

    Consider E2=mCmE_{2}=\mathbb{Z}_{m}*C_{m}. Its vertices are of the form αg(1α)g\alpha g\oplus(1-\alpha)g for gGg\in G and α{0,1}\alpha\in\{0,1\}. Similarly, we can label all rational points in mCm\|\mathbb{Z}_{m}*C_{m}\| by elements of the module [m]2\mathbb{Q}[\mathbb{Z}_{m}]^{2}. That is, x=αg(1α)bx=\alpha g\oplus(1-\alpha)b for 0α10\leq\alpha\leq 1, α\alpha\in\mathbb{Q}, gmg\in\mathbb{Z}_{m} and bb a rational point in Cm\|C_{m}\|.

  3. 3.

    For the subdivision operation, we take Δ\Delta to be the medial triangle subdivision. This process subdivides each triangle as follows:

    1. (a)

      Take a triangle {v1,v2,v3}\{v_{1},v_{2},v_{3}\}, with labels vi=aibiv_{i}=a_{i}\oplus b_{i} for ai,bi[m]a_{i},b_{i}\in\mathbb{Q}[\mathbb{Z}_{m}].

    2. (b)

      Let vij=vi+vj2=ai+aj2bi+bj2v_{ij}=\frac{v_{i}+v_{j}}{2}=\frac{a_{i}+a_{j}}{2}\oplus\frac{b_{i}+b_{j}}{2}.

    3. (c)

      Then, return the 4 triangles {v1,v12,v13}\{v_{1},v_{12},v_{13}\}, {v2,v12,v23}\{v_{2},v_{12},v_{23}\}, {v3,v13,v23}\{v_{3},v_{13},v_{23}\} and {v12,v23,v13}\{v_{12},v_{23},v_{13}\}.

    Notice that this process still produces a triangulation where all the vertices are rational points of mC2\|\mathbb{Z}_{m}*C_{2}\|. Moreover, it clearly respects the m\mathbb{Z}_{m}-action and when restricted to Cm\|C_{m}\|, it simply halves each segment.

  4. 4.

    We then follow steps 4-9 of the general approach. In our examples for m=2,,6m=2,\dots,6, we took values of k=3k=3 or 44.

Since ZmCm\|Z_{m}*C_{m}\| can be seen as mm independent disks identified by their boundaries, we can plot the vertices of TT into mm unit disks with centers at the points (Rcos(2πk/m),Rsin(2πk/m))(R\cos(2\pi k/m),R\sin(2\pi k/m)) for some large radius RR. Then, we color a fine mesh of points in the disks with the color of their closest vertex to produce Figure 1(d).

3D Case for 𝟑\bm{\mathbb{Z}_{3}}

So far, for the 3-dimensional case we have only been able to compute cov3(3)\text{cov}_{\mathbb{Z}_{3}}\left(3\right). For this we use the E23E_{2}\mathbb{Z}_{3} space and the 2\mathbb{Z}_{2}-cover described before. Just as before, we can think of labeling all its rational points by elements of the module [m]\mathbb{Q}[\mathbb{Z}_{m}]. We then apply the general approach to find a 3\mathbb{Z}_{3}-cover of E33=3E23=(32)C3E_{3}\mathbb{Z}_{3}=\mathbb{Z}_{3}*E_{2}\mathbb{Z}_{3}=(\mathbb{Z}_{3}^{*2})*C_{3}, by choosing an appropriate subdivision procedure. For this, we subdivide each tetrahedron into 4 medial tetrahedra and one central octahedron, that we further subdivide into 8 tetrahedra. That is, we proceed as follows:

  1. 1.

    Take a tetrahedron {v1,v2,v3,v4}\{v_{1},v_{2},v_{3},v_{4}\}.

  2. 2.

    For each pair of indices 1i<j41\leq i<j\leq 4, define the point vij=vi+vj2v_{ij}=\frac{v_{i}+v_{j}}{2}.

    Define the point w=v1+v2+v3+v44w=\frac{v_{1}+v_{2}+v_{3}+v_{4}}{4}.

  3. 3.

    Then, return the 12 tetrahedra:

    {v1,v12,v13,v14},\displaystyle\{v_{1},v_{12},v_{13},v_{14}\}, {v2,v12,v23,v24},{v3,v13,v23,v34},{v4,v14,v24,v34}\displaystyle\{v_{2},v_{12},v_{23},v_{24}\},\{v_{3},v_{13},v_{23},v_{34}\},\{v_{4},v_{14},v_{24},v_{34}\}
    {v12,v13,v14,w},\displaystyle\{v_{12},v_{13},v_{14},w\}, {v12,v23,v24,w},{v13,v23,v34,w},{v14,v24,v34,w},\displaystyle\{v_{12},v_{23},v_{24},w\},\{v_{13},v_{23},v_{34},w\},\{v_{14},v_{24},v_{34},w\},
    {v12,v14,v24,w},\displaystyle\{v_{12},v_{14},v_{24},w\}, {v12,v13,v23,w},{v13,v14,v34,w},{v23,v24,v34,w}\displaystyle\{v_{12},v_{13},v_{23},w\},\{v_{13},v_{14},v_{34},w\},\{v_{23},v_{24},v_{34},w\}

Notice that this procedure restricted to E23E_{2}\mathbb{Z}_{3} is just the medial triangle subdivision used in the 2-dimensional case. Moreover, ΔK\Delta K is still a 3\mathbb{Z}_{3}-simplicial complex. Thus we can finish this case by following steps 4-9 of the general approach. In our computations, we use k=3k=3, which produces a graph with 9,129 vertices, however, solving the related integer programming problem in [Gurobi] is almost immediate.

References

  • [BH99] M. R. Bridson and A. Haefliger. Mκ\kappa-polyhedral complexes. In Metric Spaces of Non-Positive Curvature, pages 97–130. Springer Berlin Heidelberg, Berlin, Heidelberg, 1999. doi:10.1007/978-3-662-12494-9.
  • [BK03] E. Babson and D. N. Kozlov. Topological obstructions to graph colorings. Electron. Res. Announc. Amer. Math. Soc., 9:61–68, 2003.
  • [BK06] E. Babson and D. N. Kozlov. Complexes of graph homomorphisms. Israel J. Math., 152:285–312, 2006.
  • [BM14] A. Barg and O. Musin, editors. Discrete Geometry and Algebraic Combinatorics, volume 625 of Contemporary Mathematics. American Mathematical Society, 2014.
  • [Dan18] H. R. Daneshpajouh. New construction of graphs with high chromatic number and small clique number. Discrete Comput. Geom., 59(1):238–245, 2018. doi:10.1007/s00454-017-9934-3.
  • [HN04] P. Hell and J. Nešetřil. Graphs and homomorphisms, volume 28 of Oxford Lecture Series in Mathematics and its Applications. Oxford University Press, Oxford, 2004.
  • [Kah07] M. Kahle. The neighborhood complex of a random graph. J. Combin. Theory Ser. A, 114(2):380–387, 2007. doi:10.1016/j.jcta.2006.05.004.
  • [Kin93] J. F. C. Kingman. Poisson Processes (Oxford Studies in Probability). Clarendon Press, jan 1993. URL https://www.xarg.org/ref/a/0198536933/.
  • [KMF20] M. Kahle and F. Martinez-Figueroa. The chromatic number of random Borsuk graphs. Random Structures Algorithms, 56(3):838–850, 2020. doi:10.1002/rsa.20897.
  • [Koz08] D. Kozlov. Combinatorial algebraic topology, volume 21 of Algorithms and Computation in Mathematics. Springer, Berlin, 2008.
  • [Lan02] S. Lang. Algebra, volume 211 of Graduate Texts in Mathematics. Springer-Verlag, New York, third edition, 2002. doi:10.1007/978-1-4613-0041-0.
  • [Lov78] L. Lovász. Kneser’s conjecture, chromatic number, and homotopy. J. Combin. Theory Ser. A, 25(3):319–324, 1978. doi:10.1016/0097-3165(78)90022-5.
  • [Mat02] J. Matoušek. Lectures on discrete geometry. Springer, New York, 2002.
  • [Mat08] J. Matoušek. Using the Borsuk–Ulam Theorem. Springer Berlin Heidelberg, 2008. doi:10.1007/978-3-540-76649-0.
  • [Pen03] M. Penrose. Random Geometric Graphs (Oxford Studies in Probability). Oxford University Press, jul 2003. URL https://www.xarg.org/ref/a/0198506260/.
  • [PRS17] R. I. Prosanov, A. M. Raĭgorodskiĭ, and A. A. Sagdeev. Improvements of the Frankl-Rödl theorem and the geometric consequences. Dokl. Akad. Nauk, 475(2):137–139, 2017. doi:10.1134/s106456241704007x.
  • [Rai12] A. M. Raigorodskii. On the chromatic numbers of spheres in n{\mathbb{R}}^{n}. Combinatorica, 32(1):111–123, 2012. doi:10.1007/s00493-012-2709-9.
  • [Sag18] A. A. Sagdeev. An improved Frankl-Rödl theorem and some of its geometric consequences. Problemy Peredachi Informatsii, 54(2):45–72, 2018. doi:10.1134/s0032946018020047.
  • [Gurobi] Gurobi Optimization, LLC. Gurobi Optimizer Reference Manual, 2021. URL https://www.gurobi.com.
  • [SageMath] The Sage Developers. SageMath, the Sage Mathematics Software System (Version 9.2), 2021. URL https://www.sagemath.org.