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

2-dimensional Shephard groups

Katherine M. Goldman
Abstract.

The 2-dimensional Shephard groups are quotients of 2-dimensional Artin groups by powers of standard generators. We show that such a quotient is not CAT(0)\mathrm{CAT}(0) if the powers taken are sufficiently large. However, for a given 2-dimensional Shephard group, we construct a CAT(0)\mathrm{CAT}(0) piecewise Euclidean cell complex with a cocompact action (analogous to the Deligne complex for an Artin group) that allows us to determine other non-positive curvature properties. Namely, we show the 2-dimensional Shephard groups are acylindrically hyperbolic (which was known for 2-dimensional Artin groups), and relatively hyperbolic (which most Artin groups are known not to be). As an application, we show that a broad class of 2-dimensional Artin groups are residually finite.

1. Introduction

Shephard groups (named for G.C. Shephard [She52]) are specific quotients of Artin groups and include, for example, the Coxeter groups and graph products of cyclic groups as special cases. In a previous paper [Gol23], the author identified a specific class of Shephard groups which exhibit Coxeter-like behavior, and proved that these are CAT(0)\mathrm{CAT}(0). However, the Shephard groups form a very broad class, and because of this, we can find examples which exhibit behavior that diverges quite a bit from the intuition that Coxeter groups and Artin groups provide. The motivating criteria in [Gol23] can be roughly summarized by saying that the finite parabolic subgroups are identical to the finite parabolic subgroups of the “associated” Coxeter group. This class is rather inflexible, though; there are only so many finite Shephard groups that are not finite Coxeter groups. Our main motivation, then, is to begin the study of Shephard groups which do not satisfy this criteria. That is to say, to study those Shephard groups which possess some infinite parabolic subgroup whose associated Coxeter group is finite.

It turns out that such groups are quite fascinating even in the 2-generator (or “dihedral”) case, and possess a rather interesting geometry which deviates from the Artin and Coxeter groups. These dihedral groups will be one of the main focuses of the current article. The other main focus are the 2-dimensional Shephard groups. These are the Shephard groups which can be reasonably said to be “built-up” out of their dihedral subgroups. From our results on 2-generator Shephard groups, we derive interesting information about the geometry of the 2-dimensional Shephard groups, some known for Artin groups, some different than Artin groups, and some which can be used to show new properties of Artin groups. The full definitions and statements are as follows.

Let Γ\Gamma be a simplicial graph with vertices ii labeled by pi2{}p_{i}\in\mathbb{Z}_{\geq 2}\cup\{\infty\} and edges {i,j}\{i,j\} labeled by mij2m_{ij}\in\mathbb{Z}_{\geq 2}. If mijm_{ij} is odd then we require pi=pjp_{i}=p_{j}. We call Γ\Gamma an extended (or Shephard) presentation graph. The Shephard group ShΓ\mathrm{Sh}_{\Gamma} with presentation graph Γ\Gamma is

ShΓ=V(Γ)|prod(i,j;mij)=prod(j,i;mij) if {i,j}E(Γ)ipi=1 if pi<,\mathrm{Sh}_{\Gamma}=\left\langle V(\Gamma)\ \middle|\ \begin{matrix}\mathrm{prod}(i,j;m_{ij})=\mathrm{prod}(j,i;m_{ij})\text{ if }\{i,j\}\in E(\Gamma)\\ i^{p_{i}}=1\text{ if }p_{i}<\infty\end{matrix}\right\rangle,

where prod(a,b;m)\mathrm{prod}(a,b;m) denotes (ab)m/2(ab)^{m/2} if mm is even and (ab)(m1)/2a(ab)^{(m-1)/2}a if mm is odd. If Γ=qpr\Gamma=\leavevmode\hbox to42.83pt{\vbox to14.68pt{\pgfpicture\makeatletter\hbox{\hskip 10.78839pt\lower-4.84348pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{{}} {}{{}}{}{{{}}{}{}{}{}{}{}{}{}}{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{1.06697pt}{0.0pt}\pgfsys@curveto{1.06697pt}{0.58928pt}{0.58928pt}{1.06697pt}{0.0pt}{1.06697pt}\pgfsys@curveto{-0.58928pt}{1.06697pt}{-1.06697pt}{0.58928pt}{-1.06697pt}{0.0pt}\pgfsys@curveto{-1.06697pt}{-0.58928pt}{-0.58928pt}{-1.06697pt}{0.0pt}{-1.06697pt}\pgfsys@curveto{0.58928pt}{-1.06697pt}{1.06697pt}{-0.58928pt}{1.06697pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } {}{{}}{} {}{}{}{}{{{}{}}}{{}}{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@lineto{21.33957pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}}{}{}{}{}{} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.75}{0.0}{0.0}{0.75}{8.8612pt}{4.10806pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$q$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{{}}{}{{{}}{}{}{}{}{}{}{}{}}{}\pgfsys@moveto{21.33957pt}{0.0pt}\pgfsys@moveto{22.40654pt}{0.0pt}\pgfsys@curveto{22.40654pt}{0.58928pt}{21.92885pt}{1.06697pt}{21.33957pt}{1.06697pt}\pgfsys@curveto{20.75029pt}{1.06697pt}{20.2726pt}{0.58928pt}{20.2726pt}{0.0pt}\pgfsys@curveto{20.2726pt}{-0.58928pt}{20.75029pt}{-1.06697pt}{21.33957pt}{-1.06697pt}\pgfsys@curveto{21.92885pt}{-1.06697pt}{22.40654pt}{-0.58928pt}{22.40654pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{21.33957pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } {{}}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}}{}{}{}{}{} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.75}{0.0}{0.0}{0.75}{-8.28864pt}{-0.8854pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$p$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}}{}{}{}{}{} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.75}{0.0}{0.0}{0.75}{25.94548pt}{-1.61458pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$r$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}} is a single edge, then we will write ShΓ=Sh(p,q,r)\mathrm{Sh}_{\Gamma}=\mathrm{Sh}(p,q,r) and call this an edge (Shephard) group or a dihedral Shephard group.

If Γ\Gamma is an extended presentation graph and p2{}p\in\mathbb{Z}_{\geq 2}\cup\{\infty\}, we denote by Γ(p)\Gamma(p) the extended presentation graph obtained from Γ\Gamma by replacing all pip_{i} with pp. In particular, for any extended presentation graph Γ\Gamma, we define

WΓ\displaystyle W_{\Gamma} =ShΓ(2)\displaystyle=\mathrm{Sh}_{\Gamma(2)}
AΓ\displaystyle A_{\Gamma} =ShΓ(),\displaystyle=\mathrm{Sh}_{\Gamma(\infty)},

which are the Coxeter group and Artin group, resp., associated to Γ\Gamma. Thus we see that every Shephard group is a quotient of an Artin group by some powers of standard generators.

Assumption.

In the rest of the paper, “extended presentation graph” will be taken to mean “extended presentation graph with all pip_{i} finite” (so the kernel of AΓShΓA_{\Gamma}\to\mathrm{Sh}_{\Gamma} contains powers of every generator), unless we explicitly mention Γ()\Gamma(\infty), and “Shephard group” will mean “Shephard group with all finite-order generators”.

We write ΛΓ\Lambda\leq\Gamma if Λ\Lambda is a full subgraph of Γ\Gamma and inherits the edge and vertex labels of Γ\Gamma. (A full or “induced” subgraph Λ\Lambda of Γ\Gamma is one where if v,wV(Λ)v,w\in V(\Lambda) and {v,w}E(Γ)\{v,w\}\in E(\Gamma), then {v,w}E(Λ)\{v,w\}\in E(\Lambda).) Then Λ\Lambda is also an extended presentation graph and thus determines a Shephard group ShΛ\mathrm{Sh}_{\Lambda}, a Coxeter group WΛW_{\Lambda}, and an Artin group AΛA_{\Lambda}. If WΛW_{\Lambda} is finite, then we will call each of Λ\Lambda, ShΛ\mathrm{Sh}_{\Lambda}, and AΛA_{\Lambda} “spherical-type”. (Similarly, if WΓW_{\Gamma} is word hyperbolic, then we will call each of Γ\Gamma, ShΓ\mathrm{Sh}_{\Gamma}, and AΓA_{\Gamma} “hyperbolic-type”.) A property of an extended presentation graph Γ\Gamma commonly of interest for Artin groups and Coxeter groups is

  1. (2D)

    For all spherical-type ΛΓ\Lambda\leq\Gamma, |V(Λ)|2|V(\Lambda)|\leq 2.

If a graph Γ\Gamma satisfies (2D), we will call Γ\Gamma 2-dimensional. Much of the behavior of 2-dimensional Shephard groups can be deduced from their dihedral subgroups. Our first main result concerns solely these dihedral Shephard groups.

Theorem A.

Let (p,q,r)(p,q,r) be a triple of integers each 2\geq 2, with p=rp=r if qq is odd, and let h=1/p+2/q+1/rh=1/p+2/q+1/r. If h1h\leq 1, then Sh(p,q,r)\mathrm{Sh}(p,q,r) cannot admit a proper action by semi-simple isometries on any CAT(0)\mathrm{CAT}(0) space. In particular, Sh(p,q,r)\mathrm{Sh}(p,q,r) is not CAT(0)\mathrm{CAT}(0). Moreover,

  1. (1)

    If h=1h=1, Sh(p,q,r)\mathrm{Sh}(p,q,r) is commensurable to the 3-dimensional integer Heisenberg group; in particular, it is virtually nilpotent and not semihyperbolic.

  2. (2)

    If h<1h<1, Sh(p,q,r)\mathrm{Sh}(p,q,r) is commensurable to the “universal central extension” of a hyperbolic surface group, is a uniform lattice in Isom(SL2~)\operatorname{Isom}(\widetilde{\mathrm{SL}_{2}\mathbb{R}}), and is biautomatic.

In particular, Sh(p,q,r)\mathrm{Sh}(p,q,r) is linear, and if an element of Sh(p,q,r)\mathrm{Sh}(p,q,r) has finite order, it is conjugate to a power of one of the standard generators.

Not mentioned is the case 1/p+2/q+1/r>11/p+2/q+1/r>1; this is because if this holds, then Sh(p,q,r)\mathrm{Sh}(p,q,r) is finite (and is handled in [Gol23]). In other words, 1/p+2/q+1/r11/p+2/q+1/r\leq 1 if and only if Sh(p,q,r)\mathrm{Sh}(p,q,r) is infinite. The fact that finite dihedral Shephard groups are linear is shown in [Cox75]; hence, the above Theorem also implies all dihedral Shephard groups are linear.

Example.

The braid group on three strands Sh(,3,)\mathrm{Sh}(\infty,3,\infty) is known to be CAT(0)\mathrm{CAT}(0) (e.g., in [BM10]), but its quotient Sh(p,3,p)\mathrm{Sh}(p,3,p) is not CAT(0)\mathrm{CAT}(0) when p6p\geq 6 (when p<6p<6, this quotient is finite).

The broadest class of Shephard groups to which Theorem A applies are described in the following.

Corollary.

Let Γ\Gamma be an extended presentation graph (possibly with infinite vertex labels) with an edge e={i,j}e=\{i,j\} such that

  1. (1)

    the parabolic She\mathrm{Sh}_{e} generated by ee embeds111Meaning She\mathrm{Sh}_{e} is isomorphic to the subgroup V(e)\langle V(e)\rangle of ShΓ\mathrm{Sh}_{\Gamma} via the map induced by the inclusion eΓe\hookrightarrow\Gamma on generators. in ShΓ\mathrm{Sh}_{\Gamma},

  2. (2)

    the edge and vertex labels of ee are finite, and

  3. (3)

    She\mathrm{Sh}_{e} is infinite

Then ShΓ\mathrm{Sh}_{\Gamma} cannot admit a proper action on a CAT(0)\mathrm{CAT}(0) space by semi-simple isometries. In particular, ShΓ\mathrm{Sh}_{\Gamma} is not CAT(0)\mathrm{CAT}(0). Moreover, 1/pi+2/mij+1/pj=11/p_{i}+2/m_{ij}+1/p_{j}=1, then ShΓ\mathrm{Sh}_{\Gamma} is not semihyperbolic.

To rephrase this corollary, an embedded infinite edge subgroup with all labels finite is a “poison subgroup” for being CAT(0)\mathrm{CAT}(0), or even semihyperbolic in some cases. It is a conjecture that given any extended presentation graph, its edge parabolics embed, so, conjecturally, (1) is unnecessary and infinite edge groups with finite labels are always poison subgroups. But, as a consequence of the upcoming Theorem B, the edge groups of a 2-dimensional presentation graph always embed, so we can say

Corollary.

Let Γ\Gamma be a 2-dimensional extended presentation graph with an edge e={i,j}e=\{i,j\} such that pi<p_{i}<\infty, pj<p_{j}<\infty, mij<m_{ij}<\infty, and 1pi+2mij+1pj1\frac{1}{p_{i}}+\frac{2}{m_{ij}}+\frac{1}{p_{j}}\leq 1. Then ShΓ\mathrm{Sh}_{\Gamma} cannot admit a proper action by semi-simple isometries on a CAT(0)\mathrm{CAT}(0) space, and in particular is not CAT(0)\mathrm{CAT}(0). Moreover, if 1pi+2mij+1pj=1\frac{1}{p_{i}}+\frac{2}{m_{ij}}+\frac{1}{p_{j}}=1, then ShΓ\mathrm{Sh}_{\Gamma} is not semihyperbolic.

Example.

Suppose Γ\Gamma is an XXL-type presentation graph (meaning all edge labels are 5\geq 5); in particular Γ\Gamma is 2-dimensional. Then AΓA_{\Gamma} is CAT(0)\mathrm{CAT}(0) [Hae22]. However there is some NN such that the quotient ShΓ(p)\mathrm{Sh}_{\Gamma(p)} is not CAT(0)\mathrm{CAT}(0) for all pNp\geq N.

In light of these examples, Theorem A may seem perplexing. Perhaps it is less surprising when one considers that the CAT(0)\mathrm{CAT}(0) condition requires non-positive curvature at all scales, and we are able to show that (most of) these groups possess other, less restrictive non-positive curvature properties.

Theorem B.

If Γ\Gamma is any 2-dimensional extended presentation graph, then its edge groups embed, and ShΓ\mathrm{Sh}_{\Gamma} acts cocompactly on a piecewise Euclidean CAT(0)\mathrm{CAT}(0) cell complex with cell stabilizers the conjugates of ShΛ\mathrm{Sh}_{\Lambda} for spherical-type ΛΓ\Lambda\leq\Gamma.

This complex is the Shephard group analogue of the Deligne complex for Artin groups and the Davis-Moussong complex for Coxeter groups. We give the full definition and proof of Theorem B in Section 5. This allows one to adapt techniques used for 2-dimensional Artin groups to show similar properties of 2-dimensional Shephard groups. As an example, the next Theorem is based on the analogous result for 2-dimensional Artin groups [Vas22].

Theorem C.

Suppose Γ\Gamma is an extended presentation graph satisfying

  1. (1)

    |V(Γ)|3|V(\Gamma)|\geq 3,

  2. (2)

    Γ\Gamma is 2-dimensional,

  3. (3)

    ShΓ\mathrm{Sh}_{\Gamma} does not split as a direct product (i.e., it is irreducible), and

  4. (4)

    every connected component of Γ\Gamma has an edge ee such that She\mathrm{Sh}_{e} is infinite.

Then ShΓ\mathrm{Sh}_{\Gamma} is acylindrically hyperbolic.

Past adapting proofs from 2-dimensional Artin groups, the condition that the vertex labels of Γ\Gamma are finite allows us to show novel results, that were either unknown or untrue for Artin groups. Our primary example is:

Theorem D.

Suppose Γ\Gamma is a hyperbolic-type, 2-dimensional extended presentation graph, and let 𝒫={ShΛ:|V(Λ)|=2,WΛ finite,ShΛ infinite}\mathcal{P}=\{\,\mathrm{Sh}_{\Lambda}:|V(\Lambda)|=2,\ W_{\Lambda}\text{ finite},\ \mathrm{Sh}_{\Lambda}\text{ infinite}\,\}, the collection of spherical-type edges of Γ\Gamma which give rise to infinite Shephard groups. Then (ShΓ,𝒫)(\mathrm{Sh}_{\Gamma},\mathcal{P}) is a relatively hyperbolic group pair. In particular, if every edge group ShΛ\mathrm{Sh}_{\Lambda} is finite, then ShΓ\mathrm{Sh}_{\Gamma} is hyperbolic.

The fact that the stabilizers of the CAT(0)\mathrm{CAT}(0) cell complex from Theorem B are precisely the edge (and vertex) subgroups indicate that in some sense, the poison subgroups are the only obstruction to being non-positively curved. The above theorem makes this idea more rigorous.

This Theorem is notable, because Artin groups are rarely hyperbolic relative to spherical-type subgroups; see [KS04] for a discussion on why this is. However, it was shown in [CC07] that Theorem D holds for Artin groups if relative hyperbolicity is replaced with “weak relative hyperbolicity” (which we will not define here). Upgrading this to relative hyperbolicity has the following noteworthy consequences:

Corollary E.

Suppose Γ\Gamma is a hyperbolic-type 2-dimensional extended presentation graph. Then ShΓ\mathrm{Sh}_{\Gamma}

  1. (1)

    has solvable word problem,

  2. (2)

    satisfies the Tits alternative,

  3. (3)

    has finite asymptotic dimension, and

  4. (4)

    has the rapid decay property.

If, in addition, there is no edge {i,j}\{i,j\} of Γ\Gamma with 1/pi+2/mij+1/pj=11/p_{i}+2/m_{ij}+1/p_{j}=1, then ShΓ\mathrm{Sh}_{\Gamma} is biautomatic.

If we restrict ourselves slightly further, it turns out we also have the following very desirable property for a large subclass of 2-dimensional Shephard groups.

Corollary F.

Suppose Γ\Gamma is an extended presentation graph with no 3-cycle (or is “triangle-free”) and no 4-cycle whose edges are each labeled 22. Then ShΓ\mathrm{Sh}_{\Gamma} is residually finite.

These are precisely the 2-dimensional graphs Γ\Gamma such that Γ\Gamma is hyperbolic-type and “type FC” (the spherical-type subgraphs are exactly the complete subgraphs).

One of the more broadly appealing aspects of Shephard groups with finite vertex labels are their potential to be used to prove results for Artin groups. For example,

Theorem G.

Suppose Γ\Gamma is a triangle-free presentation graph with no 4-cycle with all edges labeled 22. Then AΓA_{\Gamma} is residually finite.

Residual finiteness is unknown for most Artin groups, and was previously unknown even for most 2-dimensional Artin groups. For a discussion on residual finiteness of Artin groups, including some past results see [Jan22].

1.1. Organization of paper

In Section 2 we recall the relationship between central extensions and cohomology in preparation for Theorem A. Then in Section 3 we complete the proof of this Theorem by detailing the geometry of the dihedral Shephard groups through the perspective of central extensions. In Section 4, we prove the main technical lemma which implies Theorem B, namely that the dihedral Shephard groups satisfy a “syllable length” condition similar to that enjoyed by dihedral Artin groups. Following this, we construct the complex from and complete the proof of Theorem B in Section 5. The consequences of this complex being CAT(0)\mathrm{CAT}(0) are discussed in the remaining sections: Section 6 deals with acylindrical hyperbolicity, Section 7 deals with relative hyperbolicity and its consequences, and Section 8 deals with residual finiteness of the related Artin groups.

Acknowledgements

The author would like to thank Mike Davis, Jingyin Huang, and Piotr Przytycki for their valuable discussions and input. This work is partially supported by NSF grant DMS-2402105.

2. Central Extensions

We begin by recalling some background on central extensions of groups and how they relate to cohomology. Recall that if AA is abelian, an (AA-)central extension of a group GG is a group G~\tilde{G} fitting into a short exact sequence

0{0}A{A}G~{\tilde{G}}G{G}0,{0,}

with the image of AA contained in the center of G~\tilde{G}. If G~1\tilde{G}_{1} and G~2\tilde{G}_{2} are AA-central extensions of GG, then we say they are (AA-)equivalent if there is an isomorphism f:G~1G~2f:\tilde{G}_{1}\to\tilde{G}_{2} making the diagram

G~1{\tilde{G}_{1}}0{0}A{A}G{G}0{0}G~2{\tilde{G}_{2}}f\scriptstyle{f}

commute.

Notation 2.1.

For an arbitrary group GG, we will denote its identity element by ee. For an abelian group AA, we will denote its identity element by 0, and if it is cyclic, we will let 11 denote a generator.

Let E(G,A)E(G,A) denote the set of equivalence classes of AA-central extensions of GG under AA-equivalence. It is well known that there is a natural abelian group structure on E(G,A)E(G,A), under which E(G,A)H2(G;A)E(G,A)\cong H^{2}(G;A) (for example, see [Bro82, Thm. 3.12] or [DK18, §5.9.6]). There is a natural way to view the map H2(G;A)E(G,A)H^{2}(G;A)\to E(G,A). Fix a presentation G=SRG=\langle\,S\mid R\,\rangle and let YY be a K(G,1)K(G,1) such that the 2-skeleton Y2Y^{2} is the presentation complex for the given presentation of GG. Let CiC_{i} be the free abelian group on the ii-cells of YY. Then in particular, the generators of C1C_{1} are SS and the generators of C2C_{2} are RR. Let di:CiCi1d_{i}:C_{i}\to C_{i-1} be the standard cellular boundary map, and define Zi=ker(di)Z_{i}=\mathrm{ker}(d_{i}) and Bi=im(di+1)B_{i}=\mathrm{im}(d_{i+1}), so that Hi(G)=Zi/BiH_{i}(G)=Z_{i}/B_{i}. We then dualize to obtain Ci=𝖧𝗈𝗆(Ci,A)C^{i}=\mathsf{Hom}(C_{i},A), di:CiCi+1d^{i}:C^{i}\to C^{i+1} defined by φφdi+1\varphi\mapsto\varphi\circ d_{i+1}, Zi=ker(di)Z^{i}=\mathrm{ker}(d^{i}), Bi=im(di1)B^{i}=\mathrm{im}(d^{i-1}), and Hn(G;A)=Zi/BiH^{n}(G;A)=Z^{i}/B^{i}.

Now choose a class [φ]H2(G;A)=Z2/B2[\varphi]\in H^{2}(G;A)=Z^{2}/B^{2} and a representative φZ2𝖧𝗈𝗆(C2,A)\varphi\in Z^{2}\subseteq\mathsf{Hom}(C_{2},A) (or in other words, some morphism φ\varphi from the free abelian group on RR to AA). Suppose A=TQA=\langle\,T\mid Q\,\rangle is a presentation for AA. Let R~={r1φ(r):rR}\tilde{R}=\{\,r^{-1}\varphi(r):r\in R\,\} and C={[s,t]:sS,tT}C=\{\,[s,t]:s\in S,t\in T\,\}, subsets of the free group on the disjoint union STS\amalg T. We then define a group

Gφ=STR~QC.G_{\varphi}=\langle\,S\amalg T\mid\tilde{R}\amalg Q\amalg C\,\rangle. (2.1)

It is straightforward to see that the map [φ]Gφ[\varphi]\to G_{\varphi} is a well-defined injective homomorphism from H2(G;A)H^{2}(G;A) to E(G,A)E(G,A). The construction of an inverse to this map is standard and unnecessary for our purposes, so we omit it, but it may be found in [DK18, §5.9.6]. We will call [φ][\varphi] the Euler class of GφG_{\varphi}, and denote it by e(Gφ)=[φ]e(G_{\varphi})=[\varphi].

One standard result is

Proposition 2.2.

The underlying set of GφG_{\varphi} is in bijection with A×GA\times G, but GφG_{\varphi} is isomorphic to A×GA\times G as a group if and only if [φ]=0[\varphi]=0 in H2(G;A)H^{2}(G;A).

We are interested in group-theoretic properties of central extensions induced by cohomological properties of their Euler classes. One first step in this direction is provided by this Lemma:

Lemma 2.3.

Suppose G~\tilde{G} and H~\tilde{H} are AA-central extensions of a group GG such that e(H~)=ne(G~)e(\tilde{H})=ne(\tilde{G}), where nn is not a zero divisor of the \mathbb{Z}-module AA. Then H~\tilde{H} is isomorphic to a subgroup of G~\tilde{G} with index [A:nA][A:nA].

Proof.

Let [ψ]=e(H~)[\psi]=e(\tilde{H}) and [φ]=e(G~)[\varphi]=e(\tilde{G}) with representatives ψ\psi and φ\varphi, resp., chosen such that ψ=nφ\psi=n\varphi. We will write GψH~G_{\psi}\cong\tilde{H} and GφG~G_{\varphi}\cong\tilde{G}, using the notation given above for the presentation of GψG_{\psi} and GφG_{\varphi}.

Define a map Ψ:GψGφ\Psi:G_{\psi}\to G_{\varphi} by Ψ(s)=s\Psi(s)=s for sSs\in S and Ψ(t)=nt\Psi(t)=nt for tTt\in T. To check that this is a well-defined homomorphism we need to verify that Ψ(r)=e\Psi(r)=e for all rR~QTr\in\tilde{R}\amalg Q\amalg T. This is immediate for rQr\in Q or rCr\in C, so we only need to verify this for rR~r\in\tilde{R}; that is, we must verify that Ψ(r)=φ(Ψ(r))\Psi(r)=\varphi(\Psi(r)) for all rRr\in R. But this is also immediate since

Ψ(r)\displaystyle\Psi(r) =r\displaystyle=r (rr is a word in SS)
=φ(r)\displaystyle=\varphi(r)
=nψ(r)\displaystyle=n\psi(r)
=Ψ(ψ(r)).\displaystyle=\Psi(\psi(r)). (ψ(r)\psi(r) is a word in TT)

Thus this map is a well-defined homomorphism. Since nn is not a zero divisor, it is also injective. It is clear from the definition of Ψ\Psi that its image has index [A:nA][A:nA] in GφG_{\varphi}. ∎

Now consider a group homomorphism f:HGf:H\to G. This induces in a standard way a morphism f:H2(G;A)H2(H;A)f^{*}:H^{2}(G;A)\to H^{2}(H;A) on cohomology and thus a map f:E(G,A)E(H,A)f^{*}:E(G,A)\to E(H,A) on central extensions via GφHfφG_{\varphi}\mapsto H_{f^{*}\varphi}. In addition, ff can be lifted to a morphism fφ:HfφGφf_{\varphi}:H_{f^{*}\varphi}\to G_{\varphi}. This gives the commutative diagram

Hfφ{H_{f^{*}\varphi}}Gφ{G_{\varphi}}H{H}G{G}fφ\scriptstyle{f_{\varphi}}f\scriptstyle{f}

As a consequence of the above diagram commuting, we have:

Lemma 2.4.

For all [φ]H2(G;A)[\varphi]\in H^{2}(G;A), [Gφ:fφ(Hfφ)]=[G:f(H)][G_{\varphi}:f_{\varphi}(H_{f^{*}\varphi})]=[G:f(H)]. In particular, if f(H)f(H) has finite index in GG, then fφ(Hfφ)f_{\varphi}(H_{f^{*}\varphi}) has finite index in GφG_{\varphi}.

Proof.

Denote the rightmost vertical morphism of the diagram by Φ:GφG\Phi:G_{\varphi}\to G. Then fφ(Hfφ)Φ1(f(H))f_{\varphi}(H_{f^{*}\varphi})\cong\Phi^{-1}(f(H)). Since Gφ=Φ1(G)G_{\varphi}=\Phi^{-1}(G) and Φ\Phi is a surjective morphism, it follows that [Gφ:fφ(Hfφ)]=[Φ1(G):Φ1(f(H))]=[G:f(H)][G_{\varphi}:f_{\varphi}(H_{f^{*}\varphi})]=[\Phi^{-1}(G):\Phi^{-1}(f(H))]=[G:f(H)]. ∎

It is a standard fact that if ff is injective, so is fφf_{\varphi} [DK18, Ex. 5.140.3], so in particular, if HH is a finite index subgroup of GG, then for any central extension G~\tilde{G} of GG, there is a finite index subgroup H~\tilde{H} which is a central extension of HH over the same central copy of AA.

We also know:

Proposition 2.5.

[Hat02, §3.G] If n=[G:f(H)]n=[G:f(H)], then ker(f)\mathrm{ker}(f^{*}) consists only of elements whose order divides nn.

In particular, ff^{*} is always injective on the free part of H2(G;A)H^{2}(G;A).

This leads us to:

Proposition 2.6.

Suppose GG is finitely generated and AnA\cong\mathbb{Z}^{n} for some nn. If G~\tilde{G} is an AA-central extension of GG such that e(G~)e(\tilde{G}) has infinite order, then G~\tilde{G} cannot act properly by semi-simple isometries on a CAT(0)\mathrm{CAT}(0) space. In particular G~\tilde{G} is not CAT(0)\mathrm{CAT}(0).

Proof.

We want to utilize [BH13, Thm. II.6.12], which states that if a finitely generated group KK acts by isometries on a CAT(0)\mathrm{CAT}(0) space, and if KK contains a central copy of d\mathbb{Z}^{d} which acts faithfully by hyperbolic isometries (save for the identity), then this copy of d\mathbb{Z}^{d} is virtually a direct factor of KK.

To apply this to our situation, let H~\tilde{H} be a finite index subgroup of G~\tilde{G} containing the canonical copy of AnA\cong\mathbb{Z}^{n} contained in Z(G~)Z(\tilde{G}). We claim that there is no subgroup HH^{\prime} of H~\tilde{H} such that H~H×A\tilde{H}\cong H^{\prime}\times A. In order for such a subgroup to exist, we must have AZ(H~)A\subseteq Z(\tilde{H}). But then H=H~/AH=\tilde{H}/A is a subgroup of GG, and since H~\tilde{H} has finite index in G~\tilde{G}, we know HH has finite index in GG. In particular, if ι:HG\iota:H\to G denotes the inclusion morphism, then Proposition 2.5 tells us that e(H~)=ιe(G~)e(\tilde{H})=\iota^{*}e(\tilde{G}) is nontrivial in H2(H;A)H^{2}(H;A) since e(G~)e(\tilde{G}) has infinite order. Since this class is non-trivial, H~\tilde{H} does not split as a direct product by Proposition 2.2.

By [BH13, Thm. II.6.12], we conclude that if G~\tilde{G} acts by semi-simple isometries on a CAT(0)\mathrm{CAT}(0) space XX, then the central copy of AA in G~\tilde{G} cannot act faithfully by hyperbolic isometries. If the action is not faithful, then since AnA\cong\mathbb{Z}^{n} the action cannot be proper. So assume the action is not by hyperbolic isometries, meaning AA contains some non-trivial elliptic isometry γ\gamma of XX (parabolic isometries are not semi-simple). Since AnA\cong\mathbb{Z}^{n}, we know γ\langle\gamma\rangle\cong\mathbb{Z}. Since γ\gamma is elliptic, then γ\langle\gamma\rangle\cong\mathbb{Z} fixes a point of XX, and hence does not act properly.

The fact that G~\tilde{G} cannot be CAT(0)\mathrm{CAT}(0) follows from [BH13, Prop. II.6.10(2)]. (Recall that a group is CAT(0)\mathrm{CAT}(0) if it acts properly and cocompactly by isometries on a CAT(0)\mathrm{CAT}(0) space.) ∎

2.1. Central products

For some later results, it becomes useful to use a commutative version of the amalgamated product.

Definition 2.7.

Let G1G_{1}, G2G_{2}, and ZZ be any groups equipped with injective homomorphisms θi:ZZ(Gi)\theta_{i}:Z\to Z(G_{i}). Let N={(θ1(z),θ2(z)1):zZ}G1×G2N=\{\,(\theta_{1}(z),\theta_{2}(z)^{-1}):z\in Z\,\}\leq G_{1}\times G_{2}. We define the amalgamated direct product or central product of G1G_{1} and G2G_{2} over ZZ by

G1×ZG2(G1×G2)/NG_{1}\times_{Z}G_{2}\coloneqq(G_{1}\times G_{2})/N

Note that NN is a central subgroup of G1×G2G_{1}\times G_{2}, so this construction always gives a well-defined group.

Take subgroups HiH_{i} of GiG_{i} and let N=(H1×H2)NN^{\prime}=(H_{1}\times H_{2})\cap N and Z=θ11(H1)θ21(H2)ZZ^{\prime}=\theta_{1}^{-1}(H_{1})\cap\theta_{2}^{-1}(H_{2})\subseteq Z. Note that N={(θ1(z),θ2(z)1):zZ}N^{\prime}=\{\,(\theta_{1}(z),\theta_{2}(z)^{-1}):z\in Z^{\prime}\,\}. By the isomorphism theorems,

H1×ZH2=(H1×H2)/N(H1×H2)N/N(G1×G2)/N.H_{1}\times_{Z^{\prime}}H_{2}=(H_{1}\times H_{2})/N^{\prime}\cong(H_{1}\times H_{2})N/N\leq(G_{1}\times G_{2})/N.

This demonstrates H1×ZH2H_{1}\times_{Z^{\prime}}H_{2} as a natural subgroup of G1×ZG2G_{1}\times_{Z}G_{2}. By the standard proof of the isomorphism theorems, the map that realizes this inclusion is (h1,h2)N(h1,h2)N(h_{1},h_{2})N^{\prime}\mapsto(h_{1},h_{2})N. In addition, one readily sees that if HiH_{i} is finite index in GiG_{i} (i=1,2i=1,2), then H1×ZH2H_{1}\times_{Z^{\prime}}H_{2} is finite index in G1×ZG2G_{1}\times_{Z}G_{2}.

As a special case, we can take H1=G1H_{1}=G_{1} and H2={idG2}H_{2}=\{\mathrm{id}_{G_{2}}\}, in which case Z={idZ}Z^{\prime}=\{\mathrm{id}_{Z}\} (or similarly, H1={idG1}H_{1}=\{\mathrm{id}_{G_{1}}\} and H2=G2H_{2}=G_{2}). It is clear that G1G1×Z{idG2}G_{1}\cong G_{1}\times_{Z^{\prime}}\{\mathrm{id}_{G_{2}}\} via the map g(g,idG2)g\mapsto(g,\mathrm{id}_{G_{2}}), since ZZ^{\prime} is trivial. We then apply our previous remarks to this situation to obtain

Proposition 2.8.

For any groups GiG_{i} and ZZ as above, the groups G1G_{1} and G2G_{2} embed into G1×ZG2G_{1}\times_{Z}G_{2} via the maps ϵ1:g(g,idG2)N\epsilon_{1}:g\mapsto(g,\mathrm{id}_{G_{2}})N and ϵ2:g(idG1,g)N\epsilon_{2}:g\mapsto(\mathrm{id}_{G_{1}},g)N, respectively.

Just as one has the notion of internal direct product, we may define an internal central product.

Proposition 2.9.

Suppose GG is any group with subgroups H1,H2H_{1},H_{2} such that h1h2=h2h1h_{1}h_{2}=h_{2}h_{1} for all hiHih_{i}\in H_{i} (i=1,2i=1,2). Let Z=H1H2Z=H_{1}\cap H_{2}. Then H1H2GH_{1}H_{2}\leq G is isomorphic to H1×ZH2H_{1}\times_{Z}H_{2}.

Proof.

First note that since the elements of H1H_{1} and the elements of H2H_{2} commute, ZZ(Hi)Z\subseteq Z(H_{i}) for i=1,2i=1,2. Let ιi:ZHi\iota_{i}:Z\to H_{i} (i=1,2i=1,2) denote the inclusion map. Let α:H1×H2G\alpha:H_{1}\times H_{2}\to G be given by (h1,h2)h1h2(h_{1},h_{2})\mapsto h_{1}h_{2}. Clearly α\alpha is a surjective homomorphism. The kernel of α\alpha is {(h1,h2):h1=h21}\{\,(h_{1},h_{2}):h_{1}=h_{2}^{-1}\,\}, or rewritten,

{(h1,h21):h1=h2,hiHi}={(ι1(h),ι2(h)1):hH1H2=Z}.\{\,(h_{1},h_{2}^{-1}):h_{1}=h_{2},h_{i}\in H_{i}\,\}=\{\,(\iota_{1}(h),\iota_{2}(h)^{-1}):h\in H_{1}\cap H_{2}=Z\,\}.

Thus

H1H2(H1×H2)/ker(α)H1×ZH2.H_{1}H_{2}\cong(H_{1}\times H_{2})/\mathrm{ker}(\alpha)\cong H_{1}\times_{Z}H_{2}.\qed

We need one final property of central products before we can continue.

Proposition 2.10.

Take GiG_{i} and ZZ as above, and identify GiG_{i} with the subgroup of G1×ZG2G_{1}\times_{Z}G_{2} as in Proposition 2.9. Then [G1×ZG2:G1]=[G2:Z][G_{1}\times_{Z}G_{2}:G_{1}]=[G_{2}:Z] (and similarly, [G1×ZG2:G2]=[G1:Z][G_{1}\times_{Z}G_{2}:G_{2}]=[G_{1}:Z]).

This follows immediately from Proposition 2.9 and standard counting results ([G1G2:G1]=[G2:G1G2]=[G2:Z][G_{1}G_{2}:G_{1}]=[G_{2}:G_{1}\cap G_{2}]=[G_{2}:Z]).

3. Dihedral Shephard groups

The goal of this section is to study the geometry of the infinite dihedral Shephard groups in detail. We display them as \mathbb{Z}-central extensions of infinite triangle groups whose Euler class has infinite order, proving the first statement of Theorem A. Following this, we determine finer properties which we believe are interesting in their own right (and in particular complete the proof of Theorem A).

3.1. Dihedral Shephard groups as central extensions

We start by examining the second integral cohomology of the triangle groups. Let (p,q,r)(p,q,r) be a triple of positive integers, and define h=1p+1q+1rh=\frac{1}{p}+\frac{1}{q}+\frac{1}{r}. We let 𝕐=𝕊,𝔼\mathbb{Y}=\mathbb{S},\mathbb{E}, or \mathbb{H} when h>1h>1, h=1h=1, or h<1h<1, respectively. Define a triangle T(p,q,r)T(p,q,r) in 𝕐2\mathbb{Y}^{2} with vertices aa, bb, and cc, and angles π/p\pi/p, π/q\pi/q, π/r\pi/r at aa, bb, and cc, respectively. The triangle group Δ(p,q,r)\Delta(p,q,r) is the group generated by rotations of angle 2π/p2\pi/p, 2π/q2\pi/q, and 2π/r2\pi/r about the vertices aa, bb, and cc, resp., of TT in 𝕐2\mathbb{Y}^{2}. (We note that sometimes Δ(p,q,r)\Delta(p,q,r) is called a von Dyck group and “triangle group” is sometimes used to refer to 3-generator Coxeter groups.) This group has the well-known presentations

Δ(p,q,r)\displaystyle\Delta(p,q,r) =a,b,cap=bq=cr=abc=e\displaystyle=\langle\,a,b,c\mid a^{p}=b^{q}=c^{r}=abc=e\,\rangle
=a,cap=(ac)q=cr=e\displaystyle=\langle\,a,c\mid a^{p}=(ac)^{q}=c^{r}=e\,\rangle

where, by abusing notation, the generators aa, bb, and cc correspond to the respective aforementioned rotations about the vertices aa, bb, and cc of TT. We will use the second presentation to compute the cohomology. Since we’re interested in only the second integral cohomology, we describe a construction of the 3-skeleton of a K(Δ,1)K(\Delta,1).

First let K(2)K^{(2)} be the presentation complex for Δ(p,q,r)\Delta(p,q,r); that is, the cell complex with one 0-cell xx, two (oriented) 1-cells labeled aa and bb, and three 2-cells labeled eae_{a}, ece_{c}, and eace_{ac}. The attaching map of the cell eae_{a} takes the boundary ea\partial e_{a} to the loop apa^{p} with positive orientation, and similarly ece_{c} and eace_{ac} are attached to the loops crc^{r} and (ac)q(ac)^{q}, respectively.

Let K~(2)\widetilde{K}^{(2)} denote the universal cover of K(2)K^{(2)}. Note that K~(2)\widetilde{K}^{(2)} is the Cayley 2-complex of Δ(p,q,r)\Delta(p,q,r) and the 1-skeleton of K~(2)\widetilde{K}^{(2)} is the Cayley graph of Δ(p,q,r)\Delta(p,q,r). This Cayley graph is the 1-skeleton of the semiregular tiling 𝒯=𝒯(p,2q,r)\mathcal{T}=\mathcal{T}(p,2q,r) of 𝕐2\mathbb{Y}^{2} by pp-gons, 2q2q-gons, and rr-gons (e.g., [MS16]). See Figure 1 for an example. The cell structure on K~(2)\widetilde{K}^{(2)} is obtained from this tiling 𝒯\mathcal{T} by gluing (n1)(n-1) extra cells to the boundary of each nn-gon.

Refer to caption
Figure 1. The tiling 𝒯(3,6,4)\mathcal{T}(3,6,4) whose 1-skeleton is the Cayley graph of Δ(3,3,4)\Delta(3,3,4) [Kal]

We now obtain the cell complex K~\widetilde{K} by filling the generators of π2(K~(2))\pi_{2}(\widetilde{K}^{(2)}) as follows. Choose an nn-gon EE of 𝒯\mathcal{T}, and label all cells attached to E\partial E in K~(2)\widetilde{K}^{(2)} (including EE) by E1,,EnE_{1},\dots,E_{n}. The cells EiEi+1E_{i}\cup E_{i+1} (with indices mod nn) form a sphere in K~(2)\widetilde{K}^{(2)}, so attach a 3-cell Ei,i+1E_{i,i+1} to this sphere. We will denote ES=Ei,i+1E_{S}=\bigcup E_{i,i+1}. Notice that ESE_{S} is homeomorphic to a 3-sphere. We obtain K~\widetilde{K} from 𝒯\mathcal{T} by replacing each nn-gon EE with the sphere ESE_{S}. Notice that K~\widetilde{K} is the wedge of the spheres ESE_{S}. In particular, π2(K~)=0\pi_{2}(\widetilde{K})=0.

There is a natural action of Δ(p,q,r)\Delta(p,q,r) on K~\widetilde{K} coming from the action on K~(2)\widetilde{K}^{(2)} by deck transformations. For an nn-gon of 𝒯\mathcal{T}, The stabilizer of the set Ei\bigcup E_{i} is conjugate to exactly one of a\langle a\rangle, ac\langle ac\rangle, or c\langle c\rangle (depending on if n=pn=p, 2q2q, or rr, respectively). The action of this stabilizer on ESE_{S} is simply the standard action of /n\mathbb{Z}/n\mathbb{Z} on 𝕊3\mathbb{S}^{3}. This action is still free and properly discontinuous, so we may define the quotient space K=K~/ΔK=\widetilde{K}/\Delta such that K~\widetilde{K} is the universal cover of KK. Note that K(2)K^{(2)} (as defined before) is the 2-skeleton of KK.

Since K~\widetilde{K} is the universal cover of KK, we know π2(K)π2(K~)=0\pi_{2}(K)\cong\pi_{2}(\widetilde{K})=0. This means that H2(Δ)H2(K)H_{2}(\Delta)\cong H_{2}(K) and H2(Δ;)H2(K;)H^{2}(\Delta;\mathbb{Z})\cong H^{2}(K;\mathbb{Z}), where Δ=Δ(p,q,r)\Delta=\Delta(p,q,r). Note that for any nn-gon EE of 𝒯\mathcal{T}, the cell EiE_{i} of K~\widetilde{K} maps to exactly one of the cells eae_{a}, eace_{ac}, or ece_{c} if n=pn=p, 2q2q, or rr, respectively. In particular, for all such EE, the 3-cell ESE_{S} in K~\widetilde{K} maps to a single 3-cell of KK which we will denote faf_{a}, facf_{ac}, and fcf_{c} if n=pn=p, 2q2q, or rr, resp. Topologically, fg=eggx\partial f_{g}=e_{g}\cup g\cup x if g=ag=a or cc, and fac=eacacx\partial f_{ac}=e_{ac}\cup a\cup c\cup x. Moreover, the closure f¯g\overline{f}_{g} of fgf_{g} is the 33-dimensional lens space of order |g||g| with its standard cell structure for g=a,bg=a,b and f¯ac\overline{f}_{ac} is the 33-dimensional lens space of order qq with two points identified at xx.

To fix notation for the computation of the cohomology of KK, let CnC_{n} denote the free abelian group on the nn-cells of KK and dn:CnCn1d_{n}:C_{n}\to C_{n-1} the standard cellular boundary map. We then let Zn=ker(dn)Z_{n}=\mathrm{ker}(d_{n}) and Bn=im(dn+1)B_{n}=\mathrm{im}(d_{n+1}) so that Hn(K)=Zn/BnH_{n}(K)=Z_{n}/B_{n}. As usual, we dualize to obtain Cn=𝖧𝗈𝗆(Cn,)C^{n}=\mathsf{Hom}(C_{n},\mathbb{Z}), dn=(dn+1):CnCn+1d^{n}=(d_{n+1})_{*}:C^{n}\to C^{n+1} given by φφdn+1\varphi\mapsto\varphi\circ d_{n+1}, Zn=ker(dn)Z^{n}=\mathrm{ker}(d^{n}), Bn=im(dn1)B^{n}=\mathrm{im}(d^{n-1}), so that Hn(K;)=Zn/BnH^{n}(K;\mathbb{Z})=Z^{n}/B^{n}.

By a computation identical to that of the lens spaces (e.g., [Hat02, Ex. 2.43]), we see that d3d_{3} is the zero map and

d2(ea)\displaystyle d_{2}(e_{a}) =pa\displaystyle=pa
d2(ec)\displaystyle d_{2}(e_{c}) =rc\displaystyle=rc
d2(eac)\displaystyle d_{2}(e_{ac}) =q(a+c)\displaystyle=q(a+c)

This allows one to easily compute that B2=0B_{2}=0 and Z2Z_{2} (hence H2(K)H_{2}(K)) is infinite cyclic generated by

lcm(p,q,r)qeaclcm(p,r)pealcm(p,r)rec.\frac{\operatorname{lcm}(p,q,r)}{q}e_{ac}-\frac{\operatorname{lcm}(p,r)}{p}e_{a}-\frac{\operatorname{lcm}(p,r)}{r}e_{c}.
Lemma 3.1.

Define a map φC2\varphi\in C^{2} by

φ(ea)\displaystyle\varphi(e_{a}) =0\displaystyle=0
φ(ec)\displaystyle\varphi(e_{c}) =0\displaystyle=0
φ(eac)\displaystyle\varphi(e_{ac}) =1.\displaystyle=1.

Then the class [φ]H2(K,)[\varphi]\in H^{2}(K,\mathbb{Z}) has infinite order.

Proof.

Let ρ:Z2Z2/B2\rho:Z_{2}\to Z_{2}/B_{2} denote the quotient map and R:C2𝖧𝗈𝗆(Z2;)R:C^{2}\to\mathsf{Hom}(Z^{2};\mathbb{Z}) denote the restriction map ψψ|Z2\psi\mapsto\psi|_{Z_{2}}.

The restriction φ|Z2\varphi|_{Z_{2}} is nontrivial, and B2=0B_{2}=0, so ρRφ\rho\circ R\circ\varphi gives a nontrivial element of 𝖧𝗈𝗆(H2;)\mathsf{Hom}(H_{2};\mathbb{Z}). By the universal coefficient theorem, the free part of H2(K;)H^{2}(K;\mathbb{Z}) is isomorphic to 𝖧𝗈𝗆(H2;)\mathsf{Hom}(H_{2};\mathbb{Z}) via the map ψρRψ¯\psi\mapsto\rho\circ R\circ\overline{\psi} (where ψ¯\overline{\psi} is some lift of ψ\psi along the quotient map Z2Z2/B2Z^{2}\to Z^{2}/B^{2}), so it follows that the image of φ\varphi in H2(K;)H^{2}(K;\mathbb{Z}) has infinite order. ∎

Lemma 3.2.

Let φ\varphi be as above. Then Sh(p,2q,r)Δφ\mathrm{Sh}(p,2q,r)\cong\Delta_{\varphi}. In particular, Sh(p,2q,r)\mathrm{Sh}(p,2q,r) has infinite cyclic center generated by (st)q(st)^{q}.

Proof.

By the definition of Δφ\Delta_{\varphi}, we see that it has presentation

a,c,zap=cr=e,(ac)q=z,[z,a]=[z,c]=e.\langle\,a,c,z\mid a^{p}=c^{r}=e,(ac)^{q}=z,[z,a]=[z,c]=e\,\rangle.

Since zz commutes with aa and bb,

z=a1za=a1(ac)qa=(ca)q,z=a^{-1}za=a^{-1}(ac)^{q}a=(ca)^{q},

so we just as well may write

a,c,zap=cr=e,(ac)q=(ca)q=z,[z,a]=[z,c]=e.\langle\,a,c,z\mid a^{p}=c^{r}=e,(ac)^{q}=(ca)^{q}=z,[z,a]=[z,c]=e\,\rangle.

But then notice that

(ac)qa=a(ca)q=a(ac)q, and\displaystyle(ac)^{q}a=a(ca)^{q}=a(ac)^{q},\qquad\text{ and}
c(ac)q=(ca)qc=(cb)qc,\displaystyle c(ac)^{q}=(ca)^{q}c=(cb)^{q}c,

so the relations [z,a]=[z,c]=1[z,a]=[z,c]=1 are now redundant, and we may write

a,c,zap=cr=e,(ac)q=(ca)q=z.\langle\,a,c,z\mid a^{p}=c^{r}=e,(ac)^{q}=(ca)^{q}=z\,\rangle.

But now the generator zz is redundant, so we arrive at

Δφ=a,cap=cr=e,(ac)q=(ca)qSh(p,2q,r).\Delta_{\varphi}=\langle\,a,c\mid a^{p}=c^{r}=e,(ac)^{q}=(ca)^{q}\,\rangle\cong\mathrm{Sh}(p,2q,r).

Since (ac)q=zZ(Δφ)(ac)^{q}=z\in Z(\Delta_{\varphi}) and Δφ/zΔ\Delta_{\varphi}/\langle z\rangle\cong\Delta is centerless (because h1h\leq 1), it follows that (ac)q=(st)q(ac)^{q}=(st)^{q} generates the center of Sh(p,2q,r)\mathrm{Sh}(p,2q,r). ∎

Next, we examine the Shephard groups Sh(p,q,p)\mathrm{Sh}(p,q,p) for qq odd.

Lemma 3.3.

The subgroup of Sh(p,2q,2)\mathrm{Sh}(p,2q,2) generated by ss and tsttst is index-2 and isomorphic to Sh(p,q,p)\mathrm{Sh}(p,q,p).

Proof.

First, we fix the presentations of these groups as

Sh(p,q,p)\displaystyle\mathrm{Sh}(p,q,p) =σ,τσp=τp=e,στσq letters=τστq letters\displaystyle=\langle\,\sigma,\tau\mid\sigma^{p}=\tau^{p}=e,\,\underbrace{\sigma\tau\sigma\ldots}_{q\text{ letters}}=\underbrace{\tau\sigma\tau\ldots}_{q\text{ letters}}\,\rangle
Sh(p,2q,2)\displaystyle\mathrm{Sh}(p,2q,2) =s,tsp=t2=e,(st)q=(ts)q.\displaystyle=\langle\,s,t\mid s^{p}=t^{2}=e,\,(st)^{q}=(ts)^{q}\,\rangle.

Let the generator υ\upsilon of /2\mathbb{Z}/2\mathbb{Z} act on Sh(p,q,p)\mathrm{Sh}(p,q,p) by interchanging σ\sigma and τ\tau. Then the semidirect product of Sh(p,q,p)\mathrm{Sh}(p,q,p) and /2\mathbb{Z}/2\mathbb{Z} under this action has the presentation

Sh(p,q,p)/2=σ,τ,υσp=τp=e,στσq letters=τστq letters,σ=υτυ,υ2=e.\mathrm{Sh}(p,q,p)\rtimes\mathbb{Z}/2\mathbb{Z}=\langle\,\sigma,\tau,\upsilon\mid\sigma^{p}=\tau^{p}=e,\underbrace{\sigma\tau\sigma\ldots}_{q\text{ letters}}=\underbrace{\tau\sigma\tau\ldots}_{q\text{ letters}}\,,\ \sigma=\upsilon\tau\upsilon,\upsilon^{2}=e\,\rangle.

Since this is a semidirect product, the subgroup generated by σ\sigma and τ\tau is isomorphic to Sh(p,q,p)\mathrm{Sh}(p,q,p). We claim this product is isomorphic to Sh(p,2q,2)\mathrm{Sh}(p,2q,2).

The relation σ=υτυ\sigma=\upsilon\tau\upsilon makes τp=e\tau^{p}=e redundant, so we may remove it. In addition, notice that

τστq letters\displaystyle\underbrace{\tau\sigma\tau\ldots}_{q\text{ letters}} =(τσ)(q1)/2τ\displaystyle=(\tau\sigma)^{(q-1)/2}\tau
=(υσυσ)(q1)/2υσυ\displaystyle=(\upsilon\sigma\upsilon\sigma)^{(q-1)/2}\upsilon\sigma\upsilon
=(υσ)(q1)υσυ\displaystyle=(\upsilon\sigma)^{(q-1)}\upsilon\sigma\upsilon
=(υσ)qυ.\displaystyle=(\upsilon\sigma)^{q}\upsilon.

Similarly,

στσq letters\displaystyle\underbrace{\sigma\tau\sigma\ldots}_{q\text{ letters}} =(στ)(q1)/2σ\displaystyle=(\sigma\tau)^{(q-1)/2}\sigma
=(συσυ)(q1)/2σ\displaystyle=(\sigma\upsilon\sigma\upsilon)^{(q-1)/2}\sigma
=(συ)(q1)σ.\displaystyle=(\sigma\upsilon)^{(q-1)}\sigma.

Thus the relation στσq letters=τστq letters\underbrace{\sigma\tau\sigma\ldots}_{q\text{ letters}}=\underbrace{\tau\sigma\tau\ldots}_{q\text{ letters}} is equivalent to

(υσ)qυ=(συ)(q1)σ,(\upsilon\sigma)^{q}\upsilon=(\sigma\upsilon)^{(q-1)}\sigma,

which in turn is equivalent to

(υσ)q=(συ)(q1)συ=(συ)q.(\upsilon\sigma)^{q}=(\sigma\upsilon)^{(q-1)}\sigma\upsilon=(\sigma\upsilon)^{q}.

Therefore,

Sh(p,q,p)/2=σ,τ,υσp=υ2=e,(υσ)q=(συ)q,σ=υτυ.\mathrm{Sh}(p,q,p)\rtimes\mathbb{Z}/2\mathbb{Z}=\langle\,\sigma,\tau,\upsilon\mid\sigma^{p}=\upsilon^{2}=e,(\upsilon\sigma)^{q}=(\sigma\upsilon)^{q},\sigma=\upsilon\tau\upsilon\,\rangle.

And with this presentation, τ\tau is obviously redundant, so we see that

Sh(p,q,p)/2Sh(p,2q,2)\mathrm{Sh}(p,q,p)\rtimes\mathbb{Z}/2\mathbb{Z}\cong\mathrm{Sh}(p,2q,2)

via the map σs\sigma\mapsto s and υt\upsilon\mapsto t. ∎

Since Sh(p,q,p)\mathrm{Sh}(p,q,p) viewed inside Sh(p,2q,2)\mathrm{Sh}(p,2q,2) is generated by σ=s\sigma=s and τ=tst\tau=tst, we have that the element (στ)q=(st)2q(\sigma\tau)^{q}=(st)^{2q} is both in Z(Sh(p,2q,2))Z(\mathrm{Sh}(p,2q,2)) and Z(Sh(p,q,p))Z(\mathrm{Sh}(p,q,p)), since (st)2q=(stst)q=(στ)q(st)^{2q}=(stst)^{q}=(\sigma\tau)^{q}. The image of Sh(p,q,p)\mathrm{Sh}(p,q,p) under the quotient map Sh(p,2q,2)Δ(p,q,2)\mathrm{Sh}(p,2q,2)\to\Delta(p,q,2) is the subgroup DD of Δ(p,q,2)\Delta(p,q,2) generated by aa and caccac. This subgroup is finite index in Δ(p,q,2)\Delta(p,q,2) (e.g., since Sh(p,q,p)\mathrm{Sh}(p,q,p) is finite index in Sh(p,2q,2)\mathrm{Sh}(p,2q,2)). In particular,

Lemma 3.4.

Sh(p,q,p)\mathrm{Sh}(p,q,p) is a \mathbb{Z}-central extension of DD via the image of the class [φ][\varphi] from H2(Δ;)H^{2}(\Delta;\mathbb{Z}) to H2(D,)H^{2}(D,\mathbb{Z}) (which has infinite order).

Note that this restriction of [φ][\varphi] has infinite order by Proposition 2.5.

As a brief aside, using this central extension structure, we can show

Corollary 3.5.

Suppose Sh(p,q,r)\mathrm{Sh}(p,q,r) is infinite and gSh(p,q,r)g\in\mathrm{Sh}(p,q,r) has finite order. Then gg is conjugate to a power of one of the standard generators of Sh(p,q,r)\mathrm{Sh}(p,q,r).

Proof.

Let ss and tt denote the generators of order pp and rr, resp., of Sh(p,q,r)\mathrm{Sh}(p,q,r). The central quotient Sh(p,q,r)/Z\mathrm{Sh}(p,q,r)/Z of Sh(p,q,r)\mathrm{Sh}(p,q,r) acts properly and cocompactly on either 𝔼2\mathbb{E}^{2} or 2\mathbb{R}^{2}, where the generators ss and tt are sent to rotations aa and cc, resp., by an angle of 2π/p2\pi/p and 2π/r2\pi/r, resp. Since gg has finite order, it cannot be contained in the center of Sh(p,q,r)\mathrm{Sh}(p,q,r), and thus has non-trivial image g¯\overline{g} in Sh(p,q,r)/Z\mathrm{Sh}(p,q,r)/Z, still with finite order. Then by standard facts about isometries of 𝔼2\mathbb{E}^{2} and 2\mathbb{H}^{2} (and in particular, facts about the triangle groups), g¯\overline{g} is conjugate (within Sh(p,q,r)/Z\mathrm{Sh}(p,q,r)/Z) to a power of either aa or cc. Without loss of generality, we may assume g¯=h¯1akh¯\overline{g}=\overline{h}^{-1}a^{k}\overline{h} for some h¯Sh(p,q,r)/Z\overline{h}\in\mathrm{Sh}(p,q,r)/Z and 0<k<p0<k<p (the argument for conjugates of powers of cc is identical). This means there is some lift hSh(p,q,r)h\in\mathrm{Sh}(p,q,r) of h¯\overline{h} such that gZ=(h1skh)ZgZ=(h^{-1}s^{k}h)Z. In particular, there is some zZz\in Z such that g=(h1skh)z=h1(zsk)hg=(h^{-1}s^{k}h)z=h^{-1}(zs^{k})h. If zez\not=e, then zz has infinite order, thus so does zskzs^{k} (since zz commutes with sks^{k}), contradicting the assumption that gg has finite order. Thus z=ez=e and g=h1skhg=h^{-1}s^{k}h. ∎

We now prove the first part of Theorem A, namely

Theorem 3.6.

Let (p,q,r)(p,q,r) be a triple of integers each 2\geq 2, with p=rp=r if qq is odd. If 1/p+2/q+1/r11/p+2/q+1/r\leq 1, then Sh(p,q,r)\mathrm{Sh}(p,q,r) cannot admit a proper action by semi-simple isometries on any CAT(0)\mathrm{CAT}(0) space. In particular, Sh(p,q,r)\mathrm{Sh}(p,q,r) is not CAT(0)\mathrm{CAT}(0).

Proof.

First suppose qq is even, say q=2kq=2k for a positive integer kk. Then

1p+1k+1r=1p+22k+1r=1p+2q+1r1.\frac{1}{p}+\frac{1}{k}+\frac{1}{r}=\frac{1}{p}+\frac{2}{2k}+\frac{1}{r}=\frac{1}{p}+\frac{2}{q}+\frac{1}{r}\leq 1.

So by Lemma 3.2, Sh(p,q,r)=Sh(p,2k,r)\mathrm{Sh}(p,q,r)=\mathrm{Sh}(p,2k,r) is a \mathbb{Z}-central extension of the infinite group Δ(p,k,r)\Delta(p,k,r) via a second cohomology class (of infinite order by Lemma 3.1) of Δ(p,k,r)\Delta(p,k,r). Thus by Proposition 2.6, Sh(p,q,r)\mathrm{Sh}(p,q,r) cannot act properly by semi-simple isometries on a CAT(0)\mathrm{CAT}(0) space and is not CAT(0)\mathrm{CAT}(0).

Now assume qq is odd (implying p=rp=r). This means

11p+2q+1r=2p+2q,1\geq\frac{1}{p}+\frac{2}{q}+\frac{1}{r}=\frac{2}{p}+\frac{2}{q},

implying

1p+1q12,\frac{1}{p}+\frac{1}{q}\leq\frac{1}{2},

and so

1p+1q+121.\frac{1}{p}+\frac{1}{q}+\frac{1}{2}\leq 1.

Thus Sh(p,q,p)\mathrm{Sh}(p,q,p) is a \mathbb{Z}-central extension of a finite index subgroup of Δ(p,q,2)\Delta(p,q,2) via a second cohomology class of infinite order (Lemma 3.4). By Proposition 2.6 cannot act properly by semi-simple isometries on a CAT(0)\mathrm{CAT}(0) space and is not CAT(0)\mathrm{CAT}(0). ∎

3.2. Further geometry of the extension

We can give insight into the geometry of the dihedral Shephard groups beyond the general fact of Proposition 2.6. Namely, we will discuss the remainder of Theorem A.

To encompass both types of dihedral Shephard groups dealt with above (depending on the parity of qq), we fix notation for this section. First, we let Sh=Sh(p,2q,r)Sh=\mathrm{Sh}(p,2q,r) or Sh=Sh(p,q,p)Sh=\mathrm{Sh}(p,q,p) (in which case we say r=2r=2). In the first case, we define D=Δ=Δ(p,q,r)D=\Delta=\Delta(p,q,r), and in the second case, we define DD to be the subgroup of Δ=Δ(p,q,2)\Delta=\Delta(p,q,2) generated by aa and caccac. To summarize the results of the previous section in this notation, ShSh is a \mathbb{Z}-central extension of DD whose Euler class has infinite order, where DD acts geometrically on 𝔼2\mathbb{E}^{2} or 2\mathbb{H}^{2} if h=1h=1 or h<1h<1, resp., as a finite index subgroup of a triangle group Δ\Delta.

As a consequence of the latter point, DD contains a finite index torsion-free subgroup MM [Mil75, Thm. 2.7], and in particular MM must be a (closed) surface group. By Lemma 2.4, MM lifts to a finite index subgroup MφM_{\varphi} of ShSh. Since MM is a surface group, H2(M;)H^{2}(M;\mathbb{Z})\cong\mathbb{Z}. Let M~\widetilde{M} denote a central extension of MM such that e(M~)e(\widetilde{M}) is a generator of H2(M;)H^{2}(M;\mathbb{Z}). (Sometimes M~\widetilde{M} is called the “universal central extension” of MM, although this conflicts with the standard definition of universal central extension which applies only to perfect groups.) When h=1h=1, M2M\cong\mathbb{Z}^{2} and M~H(3)\widetilde{M}\cong H(3), the 3-dimensional integer Heisenberg group. When h<1h<1, then MM is a hyperbolic surface group, and M~\widetilde{M} is a uniform lattice in SL2~\widetilde{\mathrm{SL}_{2}\mathbb{R}}. Since e(M~)e(\widetilde{M}) is a generator of the second cohomology, e(Mφ)e(M_{\varphi}) is a non-zero multiple of e(M~)e(\widetilde{M}). By Lemma 2.3, this means MφM_{\varphi} is finite index in M~\widetilde{M}. Thus ShSh is commensurable to M~\widetilde{M}. As an immediate consequence, we have

Proposition 3.7.

For any triple (p,q,r)(p,q,r) of integers 2\geq 2 (with p=rp=r when qq is odd), the group Sh(p,q,r)\mathrm{Sh}(p,q,r) is linear.

Proof.

When Sh(p,q,r)\mathrm{Sh}(p,q,r) is finite, this was shown in [Cox75]. So, suppose Sh(p,q,r)\mathrm{Sh}(p,q,r) is not finite, i.e., h=1/p+2/q+1/r1h=1/p+2/q+1/r\leq 1. It is an easy exercise to see that if HH is a finite index subgroup of GG, then GG is linear if and only if HH is linear. Thus if two groups GG and HH are commensurable, one is linear if and only if the other is. So it suffices to note that M~\widetilde{M} is always linear: if h=1h=1, then M~\widetilde{M} is the 3-dimensional integer Heisenberg group H(3)H(3) (well-known to be linear), and, if h<1h<1, then M~\widetilde{M} is linear by [dLH00, §IV.48] (via an explicit injection from M~\widetilde{M} to SL2×H(3)\mathrm{SL}_{2}\mathbb{R}\times H(3)). ∎

We will note here that the Shephard group analogue of the Tits representation used to show finite Shephard groups are linear in [Cox75] is not faithful for infinite dihedral Shephard groups. (A quick computation shows that the center of the image under this representation is always finite.) Finding an explicit representation for the Shephard groups is straightforward using the information given above, so we leave it as an exercise. The h<1h<1 Shephard groups have an interesting explicit (but non-linear) representation as isometries of SL2~\widetilde{\mathrm{SL}_{2}\mathbb{R}}, which will discuss soon. First we examine the “Euclidean-like” case of h=1h=1.

Proposition 3.8.

Suppose h=1h=1. Then ShSh is virtually nilpotent and is not semihyperbolic.

Proof.

Virtual nilpotency of ShSh follows from the Q.I. rigidity of virtual nilpotency and the nilpotency of the 3-dimensional Heisenberg group. Moreover, the 3-dimensional Heisenberg group has cubic Dehn function, and thus so does ShSh. Since semihyperbolic groups have at-most quadratic Dehn function, it follows that ShSh cannot be semihyperbolic. ∎

This implies, for example, that if such a Shephard group embeds in an arbitrary Shephard group ShΓ\mathrm{Sh}_{\Gamma}, then ShΓ\mathrm{Sh}_{\Gamma} is not semihyperbolic.

The case h<1h<1 is quite rich. For example, since the central quotient is word hyperbolic in this case, by [NR97], this immediately implies

Proposition 3.9.

If h<1h<1, then ShSh is biautomatic.

Past knowing that such a Shephard group is commensurable to a uniform lattice in SL2~\widetilde{\mathrm{SL}_{2}\mathbb{R}}, we can also explicitly demonstrate it as a group of isometries of SL2~\widetilde{\mathrm{SL}_{2}\mathbb{R}}. In some sense this is “more natural” than the linearity of Proposition 3.7, because it directly generalizes the method in which the presentation for the finite dihedral Shephard groups are derived in [Cox75, §9]. (The main technical difference is the fact that 1/lcm(p,q,r)1/p+1/q+1/r11/\operatorname{lcm}(p,q,r)\not=1/p+1/q+1/r-1 when 1/p+1/q+1/r11/p+1/q+1/r\leq 1, unlike in the finite case.)

Proposition 3.10.

If h<1h<1, then ShSh is a uniform lattice in Isom(SL2~)\operatorname{Isom}(\widetilde{\mathrm{SL}_{2}\mathbb{R}}).

Proof.

Since h<1h<1 and each of pp, qq, and rr are finite, we know that Δ\Delta, DD, and MM are uniform lattices in 𝒢PSL2\mathscr{G}\coloneqq\mathrm{PSL}_{2}\mathbb{R}. Recall that 𝒢\mathscr{G} is isometric to the unit tangent bundle of 2\mathbb{H}^{2} (under the Sasaki metric), which itself can be thought of as a U(1)U(1)-bundle over 2\mathbb{H}^{2}. This bundle is topologically trivial (since 2\mathbb{H}^{2} is contractible) but is well known to be metrically non-trivial. Let 𝒢~\widetilde{\mathscr{G}} (=SL2~=\widetilde{\mathrm{SL}_{2}\mathbb{R}}) denote the universal cover of PSL2\mathrm{PSL}_{2}\mathbb{R}. Then 𝒢~\widetilde{\mathscr{G}} is a (metrically non-trivial) \mathbb{R}-bundle over 2\mathbb{H}^{2}. Note that M~\widetilde{M} (as defined above) is actually the preimage of MM under the covering map 𝒢~𝒢\widetilde{\mathscr{G}}\to\mathscr{G}. Let Δ~\widetilde{\Delta} denote the preimage of Δ\Delta under this covering map. By [Mil75, Lem. 3.1] this group has the presentation

Δ~=a~,b~,c~a~p=b~q=c~r=a~b~c~,\widetilde{\Delta}=\langle\,\tilde{a},\tilde{b},\tilde{c}\mid\tilde{a}^{p}=\tilde{b}^{q}=\tilde{c}^{r}=\tilde{a}\tilde{b}\tilde{c}\,\rangle,

where each of a~\tilde{a}, b~\tilde{b}, and c~\tilde{c} are lifts of the respective rotations aa, bb, and cc to 𝒢~\widetilde{\mathscr{G}}. Moreover, the proof of said Lemma shows that a~b~c~\tilde{a}\tilde{b}\tilde{c} generates the center of 𝒢~\widetilde{\mathscr{G}} and the center of Δ~\widetilde{\Delta}. Note that Δ~\widetilde{\Delta} is also a uniform lattice in 𝒢~\widetilde{\mathscr{G}} since Δ\Delta is a uniform lattice in 𝒢\mathscr{G}. In order to display the dihedral Shephard groups as subgroups of Isom(𝒢~)\operatorname{Isom}(\widetilde{\mathscr{G}}), we introduce another class of isometries.

For θ\theta\in\mathbb{R}, define a map rθr_{\theta} which acts on 𝒢\mathscr{G} by preserving the U(1)U(1)-fiber structure over 2\mathbb{H}^{2}, such that rθr_{\theta} projects down to the identity map of 2\mathbb{H}^{2} and rotates each fiber by 2πθ2\pi\theta. Since the bundle is topologically trivial, there is no issue with the existence and well-defined-ness of this map. Moreover, it is clear that this map is an isometry for any θ\theta. We can also see that each rθr_{\theta} commutes with the action of 𝒢\mathscr{G}. The group of all rθr_{\theta} is isomorphic to U(1)U(1). Each rθr_{\theta} can be lifted to a map r~θ\tilde{r}_{\theta} of 𝒢~\widetilde{\mathscr{G}} which translates along the \mathbb{R}-fibers a common distance 2πθ2\pi\theta. Note that this action commutes with the left action of 𝒢~\widetilde{\mathscr{G}}. Since rθr_{\theta} is an isometry of GG, it follows that r~θ\tilde{r}_{\theta} is an isometry of 𝒢~\widetilde{\mathscr{G}}. Let ={r~θ:θ}\mathscr{R}=\{\,\tilde{r}_{\theta}:\theta\in\mathbb{R}\,\}. As a straightforward exercise, one may verify that \mathscr{R} along with the left-multiplication maps of 𝒢~\widetilde{\mathscr{G}} generate the entirety of Isom(𝒢~)\operatorname{Isom}(\widetilde{\mathscr{G}}). Since the elements of \mathscr{R} commute with the elements of 𝒢~\widetilde{\mathscr{G}} (and vice versa), this means Isom(𝒢~)=𝒢~\operatorname{Isom}(\widetilde{\mathscr{G}})=\mathscr{R}\widetilde{\mathscr{G}}. But note that 𝒢~={r~θ:θ}\widetilde{\mathscr{G}}\cap\mathscr{R}=\{\,\tilde{r}_{\theta}:\theta\in\mathbb{Z}\,\}\cong\mathbb{Z}; so, by Proposition 2.9, Isom(𝒢~)𝒢~×\operatorname{Isom}(\widetilde{\mathscr{G}})\cong\widetilde{\mathscr{G}}\times_{\mathbb{Z}}\mathscr{R}.

Now let k=lcm(p,q,r)k=\mathrm{lcm}(p,q,r). Define Δ~k\widetilde{\Delta}_{k} to be the subgroup of Isom(𝒢~)\operatorname{Isom}(\widetilde{\mathscr{G}}) generated by a~\tilde{a}, b~\tilde{b}, c~\tilde{c}, and zr~1/kz\coloneqq\tilde{r}_{1/k}. Since r~1/k\tilde{r}_{1/k} commutes with the left multiplication action of 𝒢~\widetilde{\mathscr{G}}, this group has the presentation

Δ~k=a~,b~,c~,za~p=b~q=c~r=a~b~c~=zk,[a~,z]=[b~,z]=[c~,z]=e.\widetilde{\Delta}_{k}=\langle\,\tilde{a},\tilde{b},\tilde{c},z\mid\tilde{a}^{p}=\tilde{b}^{q}=\tilde{c}^{r}=\tilde{a}\tilde{b}\tilde{c}=z^{k},[\tilde{a},z]=[\tilde{b},z]=[\tilde{c},z]=e\,\rangle.

Note that this is isomorphic to a central product Δ~×r~1/k\widetilde{\Delta}\times_{\mathbb{Z}}\langle\tilde{r}_{1/k}\rangle. Since Δ~\widetilde{\Delta} is a uniform lattice in 𝒢~\widetilde{\mathscr{G}} and r~1/k\langle\tilde{r}_{1/k}\rangle is a uniform lattice in \mathscr{R}, it follows easily that Δ~k\widetilde{\Delta}_{k} is a uniform lattice in 𝒢~×Isom(𝒢~)\widetilde{\mathscr{G}}\times_{\mathbb{Z}}\mathscr{R}\cong\operatorname{Isom}(\widetilde{\mathscr{G}}). We claim that Sh(p,2q,r)\mathrm{Sh}(p,2q,r) is isomorphic to a finite index subgroup of Δ~k\widetilde{\Delta}_{k}, hence is a uniform lattice in Isom(𝒢~)\operatorname{Isom}(\widetilde{\mathscr{G}}) as well. Since Sh(p,q,p)\mathrm{Sh}(p,q,p) is finite index in Sh(p,2q,2)\mathrm{Sh}(p,2q,2), the result follows for these groups as well.

Let m=kp+kq+krkm=\frac{k}{p}+\frac{k}{q}+\frac{k}{r}-k, and consider the group

G=s,t,ϕsp=tr=e,(st)q=(ts)q=ϕqm,[s,ϕ]=[t,ϕ]=e.G=\langle\,s,t,\phi\mid s^{p}=t^{r}=e,(st)^{q}=(ts)^{q}=\phi^{-qm},[s,\phi]=[t,\phi]=e\,\rangle. (3.1)

Since 1/p+1/q+1/r=h11/p+1/q+1/r=h\not=1, we know m0m\not=0. Then GG is an amalgamated direct product of Sh(p,2q,r)\mathrm{Sh}(p,2q,r) and ϕ\langle\phi\rangle\cong\mathbb{Z} along the subgroup (st)qϕqm\mathbb{Z}\cong\langle(st)^{q}\rangle\cong\langle\phi^{-qm}\rangle. By Proposition 2.8, the subgroup of GG generated by ss and tt is isomorphic to Sh(p,2q,r)\mathrm{Sh}(p,2q,r), and by Proposition 2.10, the index of this subgroup in GG is [ϕ:ϕm]=m<[\langle\phi\rangle:\langle\phi^{m}\rangle]=m<\infty. We now show that GΔ~kG\cong\widetilde{\Delta}_{k}

We start by adding a redundant generator u=(ϕmts)1u=(\phi^{m}ts)^{-1} to GG to obtain the presentation

s,t,u,ϕsp=tr=e,(st)q=(ts)q=ϕqm,[s,ϕ]=[t,ϕ]=e,u=(ϕmts)1.\displaystyle\langle\,s,t,u,\phi\mid s^{p}=t^{r}=e,(st)^{q}=(ts)^{q}=\phi^{-qm},[s,\phi]=[t,\phi]=e,u=(\phi^{m}ts)^{-1}\,\rangle.

Define Φ:Δ~kG\Phi:\widetilde{\Delta}_{k}\to G by

a~\displaystyle\tilde{a} ϕk/ps\displaystyle\mapsto\phi^{k/p}s
b~\displaystyle\tilde{b} ϕk/qu\displaystyle\mapsto\phi^{k/q}u
c~\displaystyle\tilde{c} ϕk/rt\displaystyle\mapsto\phi^{k/r}t
z\displaystyle z ϕ,\displaystyle\mapsto\phi,

then define Ψ:GΔ~k\Psi:G\to\widetilde{\Delta}_{k} by

s\displaystyle s zk/pa~\displaystyle\mapsto z^{-k/p}\tilde{a}
u\displaystyle u zk/qb~\displaystyle\mapsto z^{-k/q}\tilde{b}
t\displaystyle t zk/rc~\displaystyle\mapsto z^{-k/r}\tilde{c}
ϕ\displaystyle\phi z\displaystyle\mapsto z

We will show that Φ\Phi and Ψ\Psi define surjective homomorphisms. Once this is shown, then clearly Φ\Phi and Ψ\Psi are mutually inverse, and thus the proof of the Proposition is complete.

Starting with Φ\Phi, we must show

Φ(a~)p=Φ(b~)q=Φ(c~)r=Φ(a~)Φ(b~)Φ(c~)=Φ(z)k\displaystyle\Phi(\tilde{a})^{p}=\Phi(\tilde{b})^{q}=\Phi(\tilde{c})^{r}=\Phi(\tilde{a})\Phi(\tilde{b})\Phi(\tilde{c})=\Phi(z)^{k}

and Φ(z)\Phi(z) commutes with each of Φ(a~)\Phi(\tilde{a}), Φ(b~)\Phi(\tilde{b}), and Φ(c~)\Phi(\tilde{c}). First, since Φ(z)=ϕ\Phi(z)=\phi and ϕ\phi is in the center of GG, the latter relation holds. We verify

Φ(a~)p\displaystyle\Phi(\tilde{a})^{p} =(ϕk/ps)p\displaystyle=(\phi^{k/p}s)^{p}
=ϕksp\displaystyle=\phi^{k}s^{p}
=ϕk\displaystyle=\phi^{k}
=Φ(z)k,\displaystyle=\Phi(z)^{k},

with an identical result for Φ(c)r\Phi(c)^{r}. Next,

Φ(b~)q\displaystyle\Phi(\tilde{b})^{q} =(ϕk/qu)q\displaystyle=(\phi^{k/q}u)^{q}
=ϕkuq\displaystyle=\phi^{k}u^{q}
=ϕk(ϕmts)q\displaystyle=\phi^{k}(\phi^{m}ts)^{-q}
=ϕkϕqm(ts)q\displaystyle=\phi^{k}\phi^{-qm}(ts)^{-q}
=ϕk(ts)q(ts)q\displaystyle=\phi^{k}(ts)^{q}(ts)^{-q}
=Φ(z)k.\displaystyle=\Phi(z)^{k}.

Last,

Φ(a~)Φ(b~)Φ(c~)\displaystyle\Phi(\tilde{a})\Phi(\tilde{b})\Phi(\tilde{c}) =(ϕk/ps)(ϕk/qu)(ϕk/rt)\displaystyle=(\phi^{k/p}s)(\phi^{k/q}u)(\phi^{k/r}t)
=ϕk/p+k/q+k/rsut\displaystyle=\phi^{k/p+k/q+k/r}sut
=ϕk/p+k/q+k/rs(ϕmts)1t\displaystyle=\phi^{k/p+k/q+k/r}s(\phi^{m}ts)^{-1}t
=ϕk/p+k/q+k/rmss1t1t\displaystyle=\phi^{k/p+k/q+k/r-m}ss^{-1}t^{-1}t
=ϕk/p+k/q+k/r(k/p+k/q+k/rk)\displaystyle=\phi^{k/p+k/q+k/r-(k/p+k/q+k/r-k)}
=ϕk\displaystyle=\phi^{k}
=Φ(z)k.\displaystyle=\Phi(z)^{k}.

In order to show Ψ\Psi is a surjective morphism, we must show

Ψ(s)p=Ψ(t)r=e,\Psi(s)^{p}=\Psi(t)^{r}=e,
(Ψ(s)Ψ(t))q=(Ψ(t)Ψ(s))q=Ψ(ϕ)qm,(\Psi(s)\Psi(t))^{q}=(\Psi(t)\Psi(s))^{q}=\Psi(\phi)^{-qm},
Ψ(u)=(Ψ(ϕ)mΨ(t)Ψ(s))1,\Psi(u)=(\Psi(\phi)^{m}\Psi(t)\Psi(s))^{-1},

and Ψ(ϕ)\Psi(\phi) commutes with Ψ(s)\Psi(s) and Ψ(t)\Psi(t). Since Ψ(ϕ)=z\Psi(\phi)=z, which is in the center of Δ~k\widetilde{\Delta}_{k}, this last relation is immediate. We begin by computing

Ψ(s)p\displaystyle\Psi(s)^{p} =(zk/pa~)p\displaystyle=(z^{-k/p}\tilde{a})^{p}
=zka~p\displaystyle=z^{-k}\tilde{a}^{p}
=zkzk\displaystyle=z^{-k}z^{k}
=e,\displaystyle=e,

with an identical computation for Ψ(t)r\Psi(t)^{r}. Before proceeding, we need a lemma regarding the relations in Δ~k\widetilde{\Delta}_{k}:

Lemma.

The relations a~b~c~=b~c~a~=c~a~b~\tilde{a}\tilde{b}\tilde{c}=\tilde{b}\tilde{c}\tilde{a}=\tilde{c}\tilde{a}\tilde{b} hold in Δ~k\widetilde{\Delta}_{k}.

Proof (of Lemma).

Since a~b~c~=zk\tilde{a}\tilde{b}\tilde{c}=z^{k} and zz is in the center of Δ~k\widetilde{\Delta}_{k}, we have

b~=a~1zkc~1=zka~1c~1=a~1c~1zk.\displaystyle\tilde{b}=\tilde{a}^{-1}z^{k}\tilde{c}^{-1}=z^{k}\tilde{a}^{-1}\tilde{c}^{-1}=\tilde{a}^{-1}\tilde{c}^{-1}z^{k}.

Solving each equation for zkz^{k} gives zk=a~b~c~=b~c~a~=c~a~b~z^{k}=\tilde{a}\tilde{b}\tilde{c}=\tilde{b}\tilde{c}\tilde{a}=\tilde{c}\tilde{a}\tilde{b}. ∎

For notational convenience, let r1=Ψ(s)r_{1}=\Psi(s), r2=Ψ(u)r_{2}=\Psi(u), and r3=Ψ(t)r_{3}=\Psi(t). The above lemma implies that r1r2r3=r2r3r1=r3r1r2r_{1}r_{2}r_{3}=r_{2}r_{3}r_{1}=r_{3}r_{1}r_{2}, and in particular r2r_{2} commutes with the product r3r1r_{3}r_{1}. Moreover, we know r2r_{2} has order qq, since (zk/qb~)q=zkb~q=zkzk=1(z^{-k/q}\tilde{b})^{q}=z^{-k}\tilde{b}^{q}=z^{-k}z^{k}=1. Now,

(Ψ(s)Ψ(t))q\displaystyle(\Psi(s)\Psi(t))^{q} =(r1r3)q\displaystyle=(r_{1}r_{3})^{q}
=r1(r3r1)qr11\displaystyle=r_{1}(r_{3}r_{1})^{q}r_{1}^{-1}
=r1r2q(r3r1)qr11\displaystyle=r_{1}r_{2}^{q}(r_{3}r_{1})^{q}r_{1}^{-1}
=r1(r2r3r1)qr11\displaystyle=r_{1}(r_{2}r_{3}r_{1})^{q}r_{1}^{-1}
=(r1r2r3)q\displaystyle=(r_{1}r_{2}r_{3})^{q} (\ast)
=(r2r3r1)q\displaystyle=(r_{2}r_{3}r_{1})^{q}
=r2q(r3r1)q\displaystyle=r_{2}^{q}(r_{3}r_{1})^{q}
=(r3r1)q\displaystyle=(r_{3}r_{1})^{q}
=(Ψ(t)Ψ(s))q\displaystyle=(\Psi(t)\Psi(s))^{q}

Moreover, using (\ast) we compute

(Ψ(s)Ψ(t))q\displaystyle(\Psi(s)\Psi(t))^{q} =(r1r2r3)q\displaystyle=(r_{1}r_{2}r_{3})^{q}
=(zk/pa~zk/qb~zk/rc~)q\displaystyle=(z^{-k/p}\tilde{a}z^{-k/q}\tilde{b}z^{-k/r}\tilde{c})^{q}
=(zk/pk/qk/ra~b~c~)q\displaystyle=(z^{-k/p-k/q-k/r}\tilde{a}\tilde{b}\tilde{c})^{q}
=(zk/pk/qk/rzk)q\displaystyle=(z^{-k/p-k/q-k/r}z^{k})^{q}
=(zk/pk/qk/r+k)q\displaystyle=(z^{-k/p-k/q-k/r+k})^{q}
=(zm)q\displaystyle=(z^{-m})^{q}
=zmq\displaystyle=z^{-mq}
=Ψ(ϕ)mq.\displaystyle=\Psi(\phi)^{-mq}.

Last,

(Ψ(ϕ)mΨ(t)Ψ(s))1\displaystyle(\Psi(\phi)^{m}\Psi(t)\Psi(s))^{-1} =[zm(zk/rc~)(zk/pa~)]1\displaystyle=[z^{m}(z^{-k/r}\tilde{c})(z^{-k/p}\tilde{a})]^{-1}
=zk/r+k/pma~1c~1\displaystyle=z^{k/r+k/p-m}\tilde{a}^{-1}\tilde{c}^{-1}
=zkk/qa~1c~1\displaystyle=z^{k-k/q}\tilde{a}^{-1}\tilde{c}^{-1}
=zk/qzka~1c~1\displaystyle=z^{-k/q}z^{k}\tilde{a}^{-1}\tilde{c}^{-1}
=zk/q(b~c~a~)a~1c~1\displaystyle=z^{-k/q}(\tilde{b}\tilde{c}\tilde{a})\tilde{a}^{-1}\tilde{c}^{-1}
=zk/qb~\displaystyle=z^{-k/q}\tilde{b}
=Ψ(u).\displaystyle=\Psi(u).\qed

4. The syllable length condition

We now turn our attention to proving Theorem B. To do this, we follow the overarching idea used to show that the Deligne complex for a 2-dimensional Artin group is CAT(0)\mathrm{CAT}(0) [CD95]. The first step in this process is to show that certain words have a minimal length. We make this precise now.

Definition 4.1.

Let SS be a finite set and W(S)W(S) the set of (finite) words in SS1S\cup S^{-1}. If wW(S)w\in W(S) with w=s1i1sninw=s_{1}^{i_{1}}\ldots s_{n}^{i_{n}} (sjSs_{j}\in S, iji_{j}\in\mathbb{Z}) is cyclically reduced, then we define the syllable length of ww with respect to SS to be (w)=S(w)=n\ell(w)=\ell_{S}(w)=n.

Definition 4.2.

If GG is a group with finite generating set SS, and wW(S)w\in W(S), then we denote by w¯G\overline{w}\in G the image of ww under the map induced by the obvious map sending the word s1i1snins_{1}^{i_{1}}\ldots s_{n}^{i_{n}} in W(S)W(S) to the element s1i1snins_{1}^{i_{1}}\ldots s_{n}^{i_{n}} in GG.

The main result of this section is the following proposition, which, as mentioned before, is one of the key steps in showing the analogue of the Deligne complex for Shephard groups is also CAT(0)\mathrm{CAT}(0). It is based on a result of Appel and Schupp [AS83], but requires a minor extra hypothesis in order to account for the torsion in the generators.

Proposition 4.3.

Consider the dihedral Shephard group Sh(p,q,r)\mathrm{Sh}(p,q,r) (with p=rp=r if qq is odd) on standard generating set S={s,t}S=\{s,t\} and identity element ee. Suppose wW(S)w\in W(S) has a cyclically reduced expression w=s1i1sninw=s_{1}^{i_{1}}\ldots s_{n}^{i_{n}}, with ijpsji_{j}\not\in\mathbb{Z}p_{s_{j}} (where ps=pp_{s}=p and pt=rp_{t}=r). If w¯=e\overline{w}=e, then (w)2q\ell(w)\geq 2q.

We will first establish notation and some brief lemmas.

Definition 4.4.

Let (p,q,r)(p,q,r) be a triple of integers all 2\geq 2 with p=rp=r if qq is odd. Let Sh=Sh(p,q,r)Sh=\mathrm{Sh}(p,q,r). If qq is even, let D=D(p,q,r)=Δ(p,q/2,r)D=D(p,q,r)=\Delta(p,q/2,r), σ=a\sigma=a, and τ=c\tau=c, and if qq is odd, let D=D(p,q,r)D=D(p,q,r) denote the subgroup of Δ(p,q,2)\Delta(p,q,2) generated by σ=a\sigma=a and τ=cac\tau=cac. Define a simplicial graph 𝒟=𝒟(p,q,r)\mathcal{D}=\mathcal{D}(p,q,r) whose vertices are the cosets of σ\langle\sigma\rangle and τ\langle\tau\rangle in DD, with an edge between two vertices if the cosets have non-trivial intersection.

Sometimes 𝒟\mathcal{D} is called a “(rank 2) coset geometry”. It is the quotient of the complex Θ^=Θ^(Sh(p,q,r))\widehat{\Theta}=\widehat{\Theta}(\mathrm{Sh}(p,q,r)) by the center of Sh(p,q,r)\mathrm{Sh}(p,q,r) (see Proposition 5.2).

Lemma 4.5.

For any triple (p,q,r)(p,q,r) (with p,q,r2p,q,r\geq 2 and p=rp=r if qq is odd), the complex 𝒟(p,q,r)\mathcal{D}(p,q,r) is (the 1-skeleton of) a tiling of either 𝔼2\mathbb{E}^{2} or 2\mathbb{H}^{2} by qq-gons.

Proof.

The case when qq is even is proven in [MS16], so assume qq is odd and D=a,cacΔ(p,q,2)D=\langle a,cac\rangle\leq\Delta(p,q,2). Note that the result holds for 𝒟(p,2q,2)\mathcal{D}(p,2q,2) (since 2q2q is even and D(p,2q,2)=Δ(p,q,2)D(p,2q,2)=\Delta(p,q,2)) and D(p,q,r)D(p,q,r) is a finite index subgroup of D(p,2q,2)D(p,2q,2) (by definition). In particular, 𝒟(p,2q,2)\mathcal{D}(p,2q,2) is a subdivision of 𝒟(p,q,p)\mathcal{D}(p,q,p); the added vertices come from adding cosets of cc, which correspond to midpoints of edges of 𝒟(p,q,p)\mathcal{D}(p,q,p). In particular, since 𝒟(p,2q,2)\mathcal{D}(p,2q,2) consists of 2q2q-gons and is the first barycentric subdivision of 𝒟(p,q,p)\mathcal{D}(p,q,p), it follows that 𝒟(p,q,p)\mathcal{D}(p,q,p) is a tiling by qq-gons. ∎

Refer to caption

(a)

Refer to caption

(b)

Figure 2. The Cayley graph (a) and coset geometry (b) for Δ(3,3,3)\Delta(3,3,3)

The graph 𝒟(p,q,r)\mathcal{D}(p,q,r) can be thought of as “collapsing” the polygons in the Cayley graph of Δ\Delta corresponding to the conjugates of the subgroups s\langle s\rangle and t\langle t\rangle. For example, the Cayley graph and coset geometry of Δ(3,3,3)\Delta(3,3,3) (coming from Sh(3,6,3)\mathrm{Sh}(3,6,3)) are shown in Figure 2. They are overlaid in Figure 3 to demonstrate how the triangles induced by the orbit of ss and tt can be shrunken to their respective cosets (where these cosets are given by solid and empty vertices, respectively).

Refer to caption
Figure 3. The Cayley graph (dashed) and coset geometry (solid) of Δ(3,3,3)\Delta(3,3,3) overlaid

We may now prove the main Proposition of this section. The argument is based on one given in [Cri05, Lemma 39].

Proof (of Prop. 4.3).

Let EE denote the edge of Θ^\widehat{\Theta} coming from the intersection of the cosets s\langle s\rangle and t\langle t\rangle in ShSh. The word ww gives rise to a path γ\gamma in Θ^\widehat{\Theta} which is the concatenation E1E2EnE_{1}E_{2}\cdots E_{n} of the edges EjE_{j} given by Ej=s1i1s2i2sj1ij1Ej1.E_{j}=s_{1}^{i_{1}}s_{2}^{i_{2}}\cdots s_{j-1}^{i_{j-1}}E_{j-1}. Since iji_{j} is not a multiple of psjp_{s_{j}}, every pair of consecutive edges in this list are distinct. In addition, notice that En=w¯EE_{n}=\overline{w}E. So, γ\gamma is a locally embedded closed loop, and in particular, the edge length (γ)=n\ell(\gamma)=n of γ\gamma is precisely the syllable length (w)\ell(w) of ww. Without loss of generality, we may assume that this loop is embedded; otherwise, we may repeat the argument on embedded subloops. In addition, we may assume without loss of generality that 0<ij<psj0<i_{j}<p_{s_{j}} by replacing iji_{j} with its remainder after division by psjp_{s_{j}}; clearly this gives the same word in ShSh and same path in Θ^\widehat{\Theta}. In particular, this assumption does not change the syllable length.

Let γ¯\overline{\gamma} be the image of γ\gamma under the covering map Θ^𝒟\widehat{\Theta}\to\mathcal{D} induced by the central quotient. This is still a closed loop in 𝒟\mathcal{D}, but now may no longer be embedded. However, can find a subpath of γ¯\overline{\gamma} which is an embedded closed loop. After reparameterization, we write γ¯\overline{\gamma} as the concatenation γ¯0γ¯1\overline{\gamma}_{0}\overline{\gamma}_{1}, with γ¯0\overline{\gamma}_{0} an embedded closed loop in 𝒟\mathcal{D} and γ¯1\overline{\gamma}_{1} not necessarily embedded, possibly trivial. Let γ0\gamma_{0} be the lift of γ¯0\overline{\gamma}_{0} to Θ^\widehat{\Theta} contained in γ\gamma and let γ1\gamma_{1} be the (possibly trivial) path in Θ^\widehat{\Theta} such that γ=γ0γ1\gamma=\gamma_{0}\gamma_{1}. Then γ0\gamma_{0} and γ1\gamma_{1} represent subwords of ww of the form w0=s1i1s2i2sjijw_{0}=s_{1}^{i_{1}}s_{2}^{i_{2}}\cdots s_{j}^{i_{j}} and w1=sj+1ij+1sj+2ij+2sninw_{1}=s_{j+1}^{i_{j+1}}s_{j+2}^{i_{j+2}}\cdots s_{n}^{i_{n}} for some 1jn1\leq j\leq n. In particular, (w)=(γ)=(γ0)+(γ1)\ell(w)=\ell(\gamma)=\ell(\gamma_{0})+\ell(\gamma_{1}).

Since γ¯0\overline{\gamma}_{0} is a non-trivial embedded loop in 𝒟\mathcal{D}, it must enclose at least one qq-gon, implying (γ0)=(γ¯0)q\ell(\gamma_{0})=\ell(\overline{\gamma}_{0})\geq q. We claim that γ1\gamma_{1} is non-empty, or in other words, that γγ0\gamma\not=\gamma_{0}. Since γ\gamma is closed, it suffices to show that γ0\gamma_{0} is not a closed loop. If we show this, then, since γ¯1\overline{\gamma}_{1} will be a (nontrivial) closed path in 𝒟\mathcal{D}, we can apply the argument given for γ¯\overline{\gamma} and γ¯0\overline{\gamma}_{0} to γ¯1\overline{\gamma}_{1} and a simple subpath of γ¯1\overline{\gamma}_{1} to show that (γ¯1)q\ell(\overline{\gamma}_{1})\geq q, and thus (w)(γ¯0)+(γ¯1)q+q=2q\ell(w)\geq\ell(\overline{\gamma}_{0})+\ell(\overline{\gamma}_{1})\geq q+q=2q, as claimed.

Showing that γ0\gamma_{0} is not closed is equivalent to showing w0¯e\overline{w_{0}}\not=e. If qq is even, let Δ=Δ(p,q/2,r)\Delta=\Delta(p,q/2,r) and if qq is odd, let Δ=Δ(p,q,2)\Delta=\Delta(p,q,2) (so either D=ΔD=\Delta or DD is finite index in Δ\Delta). Let KK be the (3-skeleton of a) K(Δ,1)K(\Delta,1) space defined in Section 3, with universal cover K~\widetilde{K}, and let CC be the Cayley graph of ShSh. Note that CC is a covering of K~(1)\widetilde{K}^{(1)} (with K~(1)\widetilde{K}^{(1)} the Cayley graph of Δ\Delta). The word w0w_{0} gives rise to a path in CC in the standard way, hence also a path ρ~\tilde{\rho} in K~(1)\widetilde{K}^{(1)} via the covering map CK~(1)C\to\widetilde{K}^{(1)}, and a path ρK(1)\rho\in K^{(1)} under the covering map K~K\widetilde{K}\to K. The path ρ\rho induces a cycle ρ¯C1\overline{\rho}\in C_{1} (see Section 3 for notation). Since w0w_{0} represents the trivial word in Δ\Delta, [ρ¯]=0[\overline{\rho}]=0. This means ρ¯B1=im(d2)\overline{\rho}\in B_{1}=\mathrm{im}(d_{2}), so we may choose an element RC2R\in C_{2} such that d2(R)=ρ¯d_{2}(R)=\overline{\rho}, say R=naea+ncec+naceacR=n_{a}e_{a}+n_{c}e_{c}+n_{ac}e_{ac} for some na,nc,nacn_{a},n_{c},n_{ac}\in\mathbb{Z}. Since γ0\gamma_{0} is simple and we have assumed 0<ij<psj0<i_{j}<p_{s_{j}}, we know that ρ~\tilde{\rho} is a simple loop, it must enclose at least one cell of K~(2)\widetilde{K}^{(2)} which maps to eace_{ac}, and it must traverse the boundaries of each such cell with a consistent (positive) orientation. This means that nac0n_{ac}\not=0, so φ(R)0\varphi(R)\not=0. Therefore w¯0\overline{w}_{0} cannot be trivial in Sh(p,q,r)\mathrm{Sh}(p,q,r) if qq is even. If qq is odd this shows that w¯0\overline{w}_{0} is not trivial in Sh(p,2q,2)\mathrm{Sh}(p,2q,2), but since DD lifts to Sh(p,q,p)Sh(p,2q,2)\mathrm{Sh}(p,q,p)\leq\mathrm{Sh}(p,2q,2) and w¯0\overline{w}_{0} lies in this subgroup by assumption, the result also holds for qq odd. ∎

Example 4.6.

We will give an illustrative example with the complexes 𝒟(p,q,r)\mathcal{D}(p,q,r) and Θ^\widehat{\Theta}. (We use these over the Cayley graphs since the figures are much clearer, but a similar conceptualization works for the Cayley graphs.)

Refer to caption

(a)

Refer to caption

(b)

Figure 4. Hexagons in 𝒟(3,3,3)\mathcal{D}(3,3,3) and a path enclosing them

Consider the two hexagons in 𝒟(3,3,3)\mathcal{D}(3,3,3) highlighted in Figure 4(a). Figure 5(a) shows lifts of these hexagons to the complex Θ^\widehat{\Theta} for the Shephard group Sh(3,6,3)\mathrm{Sh}(3,6,3). (The dashed lines show which vertices are identified under the covering map; they are not part of Θ^\widehat{\Theta}.)

Refer to caption

(a)

Refer to caption

(b)

Figure 5. Lifts to Θ^\widehat{\Theta} for Sh(3,6,3)\mathrm{Sh}(3,6,3)

The entire preimage of a hexagon under the covering map is a vertical column which resembles the universal cover of the circle with the cell structure coming from the hexagon. Consider the path encircling these two hexagons in 𝒟(3,3,3)\mathcal{D}(3,3,3), shown in red in Figure 4(b). Its lift to Θ^\widehat{\Theta} is shown in Figure 5(b). The endpoints of this path are “distance 2” along the fiber of the base vertex. This corresponds to the fact that the path encloses exactly two hexagons. One may compare this to the usual description of the Cayley graph of the 3-dimensional integer Heisenberg group H(3)H(3), the main difference being that the vertical dashed lines would be actual edges of the Cayley graph of H(3)H(3). This also illustrates how the center of Sh(p,q,r)\mathrm{Sh}(p,q,r) acts on its complex Θ^\widehat{\Theta}, since it acts by deck transformations; it is a uniform “vertical translation” along the dashed lines in Figure 5 (the fibers of the map to 𝒟(p,q,r)\mathcal{D}(p,q,r)).

5. A CAT(0)\mathrm{CAT}(0) cell complex for 2-dimensional Shephard groups

In this section, we recall the definition of Θ(Γ)\Theta(\Gamma) for an arbitrary extended presentation graph Γ\Gamma, which largely follows [Gol23, §3]. We then show that this complex is CAT(0)\mathrm{CAT}(0) when Γ\Gamma is 2-dimensional.

Definition 5.1.

Let Γ\Gamma be an extended presentation graph. We define K=KΓ=|(𝒮Γf)|K=K_{\Gamma}=|(\mathcal{S}^{f}_{\Gamma})^{\prime}|, where (𝒮Γf)(\mathcal{S}^{f}_{\Gamma})^{\prime} denotes the derived complex of 𝒮Γf\mathcal{S}^{f}_{\Gamma} and |(𝒮Γf)||(\mathcal{S}^{f}_{\Gamma})^{\prime}| is its geometric realization. We will denote an nn-simplex of KK by

[Λ0<Λ1<<Λn][\Lambda_{0}<\Lambda_{1}<\dots<\Lambda_{n}]

for a chain Λ0<Λ1<<Λn\Lambda_{0}<\Lambda_{1}<\dots<\Lambda_{n} with each Λi𝒮f\Lambda_{i}\in\mathcal{S}^{f}. We note that the vertices are indexed by elements of 𝒮f\mathcal{S}^{f}; we will let vΛ=[Λ]v_{\Lambda}=[\Lambda] denote the vertex of KK coming from Λ\Lambda. Define a complex of groups 𝒢=𝒢(ShΓ,KΓ)\mathcal{G}=\mathcal{G}(\mathrm{Sh}_{\Gamma},K_{\Gamma}) over KK by declaring the local group at vΛv_{\Lambda} to be ShΛ\mathrm{Sh}_{\Lambda} and the edge maps to be the natural maps coming from the inclusion of generators.

Next, define Δ=ΔΓ\Delta=\Delta_{\Gamma} to be a simplex whose vertices are labeled by the generators V(Γ)V(\Gamma) of ShΓ\mathrm{Sh}_{\Gamma}. For ΛΓ\Lambda\subseteq\Gamma, let σΛ\sigma_{\Lambda} denote the face of ΔΓ\Delta_{\Gamma} spanned by the elements of V(Λ)V(\Lambda). We define a complex of groups 𝒢^=𝒢^(ShΓ,ΔΓ)\widehat{\mathcal{G}}=\widehat{\mathcal{G}}(\mathrm{Sh}_{\Gamma},\Delta_{\Gamma}) by declaring the local group at the face σΛ\sigma_{\Lambda} to be the group ShΛ^\mathrm{Sh}_{\widehat{\Lambda}}, where Λ^\widehat{\Lambda} is the full subgraph of Γ\Gamma generated by the vertices V(Γ)V(Λ)V(\Gamma)\setminus V(\Lambda). The edge maps are the standard maps induced by the inclusion of generating sets.

Note that 𝒢\mathcal{G} is a simple complex of groups. Hence by [BH13, Def. II.12.12], the fundamental group of 𝒢\mathcal{G} is the direct limit over the edge maps; when Γ\Gamma is 2-dimensional, this direct limit is clearly ShΓ\mathrm{Sh}_{\Gamma}. In general, neither complex of groups is a priori developable. If 𝒢\mathcal{G} is developable, we will denote its development Θ=Θ(Γ)=ΘΓ\Theta=\Theta(\Gamma)=\Theta_{\Gamma}. If 𝒢^\widehat{\mathcal{G}} is developable, we will denote its development Θ^=Θ^(Γ)=Θ^Γ\widehat{\Theta}=\widehat{\Theta}(\Gamma)=\widehat{\Theta}_{\Gamma}. For dihedral Shephard groups, it turns out that Θ^\widehat{\Theta} has a straightforward description.

Proposition 5.2.

If Λ\Lambda is the graph which is a single edge between vertices ss and tt with all labels finite, then 𝒢^(ShΛ,ΔΛ)\widehat{\mathcal{G}}(\mathrm{Sh}_{\Lambda},\Delta_{\Lambda}) is developable, and Θ^Λ\widehat{\Theta}_{\Lambda} can be described as follows: the vertex set of Θ^Λ\widehat{\Theta}_{\Lambda} are the cosets of s\langle s\rangle and t\langle t\rangle in ShΛ\mathrm{Sh}_{\Lambda}, and there is an edge between two vertices if the corresponding cosets intersect nontrivially. In particular, the center acts freely on Θ^Λ\widehat{\Theta}_{\Lambda}.

Proof.

The fact that 𝒢^\widehat{\mathcal{G}} is developable follows from the characterizations of Sh(p,q,r)\mathrm{Sh}(p,q,r) given in Section 3. Namely, in each case the cyclic groups /p\mathbb{Z}/p\mathbb{Z} and /r\mathbb{Z}/r\mathbb{Z} embed as the vertex groups. The statement regarding the vertices and edges of Θ^\widehat{\Theta} is identical to Coxeter groups and Artin groups; e.g., [CD95, Proof of Lemma 4.3.2]. The center acts freely because the stabilizers of vertices and edges are finite (they are conjugates of s\langle s\rangle or t\langle t\rangle in the first case, and are trivial in the second), while the center has infinite order. ∎

When 𝒢\mathcal{G} is developable, it too has a straightforward description.

Proposition 5.3.

Let ShΓ𝒮Γf={gShΛ:Λ𝒮Γf}\mathrm{Sh}_{\Gamma}\mathcal{S}^{f}_{\Gamma}=\{\,g\mathrm{Sh}_{\Lambda}:\Lambda\in\mathcal{S}^{f}_{\Gamma}\,\}, ordered by inclusion. If 𝒢(ShΓ,KΓ)\mathcal{G}(\mathrm{Sh}_{\Gamma},K_{\Gamma}) is developable, then ΘΓ|(ShΓ𝒮Γf)|\Theta_{\Gamma}\cong|(\mathrm{Sh}_{\Gamma}\mathcal{S}^{f}_{\Gamma})^{\prime}|. In particular, the nn-simplices of ΘΓ\Theta_{\Gamma} correspond to (n+1)(n+1)-chains g0ShΛ0<<gnShΛng_{0}\mathrm{Sh}_{\Lambda_{0}}<\dots<g_{n}\mathrm{Sh}_{\Lambda_{n}} of cosets, and the action of ShΓ\mathrm{Sh}_{\Gamma} is just the left multiplication action. We will denote the simplex arising from such a chain by

[g0ShΛ0<<gnShΛn].[g_{0}\mathrm{Sh}_{\Lambda_{0}}<\dots<g_{n}\mathrm{Sh}_{\Lambda_{n}}].

The stabilizer of this simplex is g01ShΛ0g0g_{0}^{-1}\mathrm{Sh}_{\Lambda_{0}}g_{0}.

With this description, for developable 𝒢\mathcal{G} it is clear that KK embeds in ΘΓ\Theta_{\Gamma} as the subcomplex consisting of simplices of the form [ShΛ0,ShΛ1,,ShΛn][\mathrm{Sh}_{\Lambda_{0}},\mathrm{Sh}_{\Lambda_{1}},\dots,\mathrm{Sh}_{\Lambda_{n}}]. This allows one to easily see that KK is a fundamental domain for the action of ShΛ\mathrm{Sh}_{\Lambda} on ΘΛ\Theta_{\Lambda}. In particular, we may view the vertices vΛv_{\Lambda} as being in ΘΓ\Theta_{\Gamma} as well as KK.

In order to show 𝒢\mathcal{G} is developable (under certain conditions), we show that it is “nonpositively curved” in the following sense.

Lemma 5.4.

[BH13, Thm. II.12.28] If \mathcal{H} is a (simple) complex of groups over a simply connected domain and the local development at each vertex is locally CAT(0)\mathrm{CAT}(0), then \mathcal{H} is developable and has locally CAT(0)\mathrm{CAT}(0) development.

It is clear that KK is simply connected, since vv_{\varnothing} is a cone point. We will describe the local development shortly. We begin with putting a metric on the fundamental domain KK.

Definition 5.5.

For Λ𝒮f\Lambda\in\mathcal{S}^{f}, let

𝒮Λf\displaystyle\mathcal{S}^{f}_{\geq\Lambda} ={Λ𝒮f:ΛΛ}\displaystyle=\{\,\Lambda^{\prime}\in\mathcal{S}^{f}:\Lambda^{\prime}\geq\Lambda\,\} FΛ\displaystyle F_{\Lambda} =|(𝒮Λf)|\displaystyle=|(\mathcal{S}^{f}_{\geq\Lambda})^{\prime}|
𝒮Λf\displaystyle\mathcal{S}^{f}_{\leq\Lambda} ={Λ𝒮f:ΛΛ}\displaystyle=\{\,\Lambda^{\prime}\in\mathcal{S}^{f}:\Lambda^{\prime}\leq\Lambda\,\} FΛ\displaystyle F_{\Lambda}^{*} =|(𝒮Λf)|.\displaystyle=|(\mathcal{S}^{f}_{\leq\Lambda})^{\prime}|.

Notice that FΛF_{\Lambda}^{*} is combinatorially a cube whose faces are FΛ1FΛ2F_{\Lambda_{1}}\cap F_{\Lambda_{2}}^{*} where Λ1Λ2Λ\Lambda_{1}\subseteq\Lambda_{2}\subseteq\Lambda. So KK itself has a cubical structure with faces FΛ1FΛ2F_{\Lambda_{1}}\cap F_{\Lambda_{2}}^{*} for Λ1Λ2𝒮f\Lambda_{1}\subseteq\Lambda_{2}\in\mathcal{S}^{f}. This Proposition is a straightforward exercise:

Proposition 5.6.

With the cubical cell structure above, FΛF_{\Lambda}^{*} is isomorphic to the cone on the first barycentric subdivision ΔΛ\Delta_{\Lambda}^{\prime} of ΔΛ\Delta_{\Lambda} with cone point vΛv_{\Lambda}. The isomorphism is induced by the map sending vΛ0v_{\Lambda_{0}} to the barycenter of σΛ0^\sigma_{\widehat{\Lambda_{0}}}.

Thus we may identify lk(vΛ,FΛ)lk(v_{\Lambda},F_{\Lambda}^{*}) and ΔΛ\Delta_{\Lambda}. With this connection we can define an explicit metric on KK.

Definition 5.7.

We give the cell FΛ1FΛ2F_{\Lambda_{1}}\cap F_{\Lambda_{2}}^{*} of KK the metric of a Coxeter block. Briefly, this is (the closure of) a connected component of the Coxeter zonotope associated to the finite Coxeter group WΛ2W_{\Lambda_{2}} minus its reflection hyperplanes. (See [CD95, §4.4], where this metric is defined in detail for the Davis-Moussong complex and Deligne complex.) In this metric, if Λ𝒮f\Lambda\in\mathcal{S}^{f}, then ΔΛ\Delta_{\Lambda} is a spherical simplex where the length of the edge between the vertices corresponding to ii and jj is π/mij\pi/{m_{ij}}.

Since KK is a (strict) fundamental domain, we can use the action of ShΛ\mathrm{Sh}_{\Lambda} to metrize all of ΘΓ\Theta_{\Gamma}. We will call this the Moussong metric on ΘΓ\Theta_{\Gamma}.

Now we recall the local development, focusing on the case of 𝒢\mathcal{G}. Let vΛv_{\Lambda} be a vertex of KK, with Λ𝒮f\Lambda\in\mathcal{S}^{f}. The upper star StΛSt^{\Lambda} of vΛv_{\Lambda} in 𝒢\mathcal{G} is the (full) subcomplex of KK spanned by the vertices vΛv_{\Lambda^{\prime}} with ΛΛ\Lambda^{\prime}\geq\Lambda. The lower link LkΛLk_{\Lambda} of vΛv_{\Lambda} in 𝒢\mathcal{G} is the development of the subcomplex of groups 𝒢^(K<Λ)\widehat{\mathcal{G}}(K_{<\Lambda}) of 𝒢(K)\mathcal{G}(K), where K<ΛK_{<\Lambda} denotes the subcomplex spanned by vertices vΛv_{\Lambda^{\prime}} with ΛΛ\Lambda^{\prime}\lneq\Lambda. Both of these objects are simplicial complexes which inherit the metric placed on KK. The local development at vΛv_{\Lambda} is the spherical join

D(Λ)=StΛLkΛ.\displaystyle D(\Lambda)=St^{\Lambda}\ast Lk_{\Lambda}.

Its metric naturally comes from the metric on KK. The link of vΛv_{\Lambda} in the local development is

lk(vΛ,D(Λ))\displaystyle lk(v_{\Lambda},D(\Lambda)) =LkΛLkΛ,\displaystyle=Lk^{\Lambda}\ast Lk_{\Lambda},

where LkΛLk^{\Lambda} is the upper link, meaning the (full) subcomplex of KK spanned by the vertices vΛv_{\Lambda^{\prime}} with ΛΛ\Lambda^{\prime}\gneq\Lambda. We may also sometimes refer to this complex as K>ΛK_{>\Lambda}.

Note that K>ΛK_{>\Lambda} is isomorphic to lk(vΛ,FΛ)lk(v_{\Lambda},F_{\Lambda}) and K<ΛK_{<\Lambda} is isomorphic to lk(vΛ,FΛ)lk(v_{\Lambda},F^{*}_{\Lambda}). We use the previous proposition to identify K<ΛK_{<\Lambda} with ΔΛ\Delta_{\Lambda}. With this identification, the complex of groups 𝒢^(K<Λ)\widehat{\mathcal{G}}(K_{<\Lambda}) is isomorphic to 𝒢^(GΛ,ΔΛ)\widehat{\mathcal{G}}(G_{\Lambda},\Delta_{\Lambda}) as defined above, and thus LkΛLk_{\Lambda} is isomorphic to Θ^Λ\widehat{\Theta}_{\Lambda}. It is straightforward to check that the metrics placed on KK above agree with the claimed metrics on ΔΛ\Delta_{\Lambda}. In other words, there is an isometry

lk(vΛ,D(Λ))lk(vΛ,FΛ)Θ^Λ,\displaystyle lk(v_{\Lambda},D(\Lambda))\cong lk(v_{\Lambda},F_{\Lambda})*\widehat{\Theta}_{\Lambda},

where the join is the usual spherical join. There are two special cases to note. When Λ=\Lambda=\varnothing, then Θ^\widehat{\Theta}_{\varnothing} is empty, so lk(vΛ,D(Λ))lk(vΛ,FΛ)lk(v_{\Lambda},D(\Lambda))\cong lk(v_{\Lambda},F_{\Lambda}). When Λ\Lambda is maximal in 𝒮f\mathcal{S}^{f}, FΛF_{\Lambda} is a single point vΛv_{\Lambda}, so lk(vΛ,FΛ)lk(v_{\Lambda},F_{\Lambda}) is empty and lk(vΛ,D(Λ))Θ^Λlk(v_{\Lambda},D(\Lambda))\cong\widehat{\Theta}_{\Lambda}.

Showing the local development is nonpositively curved amounts to showing that these components of the links are CAT(1)\mathrm{CAT}(1). Since a spherical join is CAT(1)\mathrm{CAT}(1) if and only if both components are [BH13, Cor. II.3.15], this reduces to showing that lk(vΛ,FΛ)lk(v_{\Lambda},F_{\Lambda}) and Θ^Λ\widehat{\Theta}_{\Lambda} are CAT(1)\mathrm{CAT}(1) when Λ𝒮f\Lambda\in\mathcal{S}^{f}. With this in mind, we can now complete the proof of Theorem B. It will follow immediately by the next theorem and general facts about simple complexes of groups.

Theorem 5.8.

Suppose Γ\Gamma is a 2-dimensional extended presentation graph. Then 𝒢\mathcal{G} is developable and its development ΘΓ\Theta_{\Gamma} is CAT(0)\mathrm{CAT}(0).

Proof.

As in the Artin group and Coxeter group case, lk(vΛ,FΛ)lk(v_{\Lambda},F_{\Lambda}) is CAT(1)\mathrm{CAT}(1) whenever Λ𝒮Γf\Lambda\in\mathcal{S}^{f}_{\Gamma} (see discussion before and after Lemma 4.4.1 in [CD95]), so it remains to show that Θ^Λ\widehat{\Theta}_{\Lambda} is CAT(1)\mathrm{CAT}(1) for Λ𝒮Γf\Lambda\in\mathcal{S}^{f}_{\Gamma}. Since Γ\Gamma is 2-dimensional, the only elements of 𝒮f\mathcal{S}^{f} are \varnothing, singletons, and edges. In the first two cases, Θ^Λ\widehat{\Theta}_{\Lambda} is either empty or finite, resp., so we may assume we are in the third case. Note that if ShΛ\mathrm{Sh}_{\Lambda} is finite (i.e., has labels 1/pi+2/mij+1/pj>11/p_{i}+2/m_{ij}+1/p_{j}>1), then this complex was shown to be CAT(1)\mathrm{CAT}(1) in [Gol23, Lemma 6.1], so we may assume this group is infinite.

By [BH13, Lem. II.5.6], it suffices to show that Θ^Λ\widehat{\Theta}_{\Lambda} has no closed loops of length <2π<2\pi. Since the length of an edge of Θ^Λ\widehat{\Theta}_{\Lambda} is π/mij\pi/m_{ij}, we must show that the edge length of any closed loop in Θ^Λ\widehat{\Theta}_{\Lambda} is at least 2mij2m_{ij}. Let γ\gamma be an embedded closed loop in Θ^Λ\widehat{\Theta}_{\Lambda}, say γ=E1En\gamma=E_{1}\cdots E_{n} for edges EiE_{i} of Θ^\widehat{\Theta}. Note that the edge path length (γ)\ell(\gamma) is nn. For 1in1\leq i\leq n, let viv_{i} be the vertex at which the edges Ei1E_{i-1} and EiE_{i} meet (with indices taken cyclically). Let sis_{i} be the generator of ShΛ\mathrm{Sh}_{\Lambda} such that vi=gisiv_{i}=g_{i}\langle s_{i}\rangle for some giShΛg_{i}\in\mathrm{Sh}_{\Lambda}. We can write Ei=gisimiEi1E_{i}=g_{i}s_{i}^{m_{i}}E_{i-1} for some mim_{i}\in\mathbb{Z}. Since γ\gamma is embedded, mim_{i} is not a multiple of psip_{s_{i}} (the order of sis_{i}). Stringing together these equalities gives a word s1m1s2m2snmn=1s_{1}^{m_{1}}s_{2}^{m_{2}}\cdots s_{n}^{m_{n}}=1 in ShΛ\mathrm{Sh}_{\Lambda}. Since sisi+1s_{i}\not=s_{i+1}, the syllable length of this word is n=(γn=\ell(\gamma. By Proposition 4.3, we must have (γ)=n2mij\ell(\gamma)=n\geq 2m_{ij}. The result now follows from Lemma 5.4. ∎

When the local groups of a nonpositively curved complex of groups are all finite, this determines all finite subgroups of the fundamental group (namely, they are the conjugates of the local groups). While we can’t exactly say that here, we can determine all elements of finite order.

Corollary 5.9.

Suppose Γ\Gamma is a 2-dimensional extended presentation graph. If hShΓh\in\mathrm{Sh}_{\Gamma} has finite order, then it is conjugate to a power of one of the standard generators of ShΓ\mathrm{Sh}_{\Gamma}.

Proof.

Suppose hShΓ{e}h\in\mathrm{Sh}_{\Gamma}\setminus\{e\} has finite order. By [BH13, Cor. II.2.8(1)], the fixed point set Fix(h)={xΘΓ:hx=x}Fix(h)=\{\,x\in\Theta_{\Gamma}:hx=x\,\} is a non-empty convex subset of ΘΓ\Theta_{\Gamma}. Since Γ\Gamma is 2-dimensional, the simplices of ΘΓ\Theta_{\Gamma} are at most dimension 2, and these top-dimensional simplices are of the form

[g0Sh,g1Sh{s},g2She][g_{0}\mathrm{Sh}_{\varnothing},g_{1}\mathrm{Sh}_{\{s\}},g_{2}\mathrm{Sh}_{e}]

for ee an edge of Γ\Gamma and ss a vertex of ee. But the stabilizer of any point xx in the interior of such a cell is g01Shg0g_{0}^{-1}\mathrm{Sh}_{\varnothing}g_{0}, which is the trivial group. In particular, Fix(h)Fix(h) must be a tree in the 1-skeleton of ΘΓ\Theta_{\Gamma} and avoid vertices of type [gSh][g\mathrm{Sh}_{\varnothing}]. If Fix(h)Fix(h) contains a vertex V=[gSh{s}]V=[g\mathrm{Sh}_{\{s\}}] or an edge E=[gSh{s},gShe]E=[g\mathrm{Sh}_{\{s\}},g^{\prime}\mathrm{Sh}_{e}], then hStab(V)=g1Sh{s}gh\in Stab(V)=g^{-1}\mathrm{Sh}_{\{s\}}g and so is conjugate to a power of ss. So suppose neither of these cases occur. This implies Fix(h)=[gShe]Fix(h)=[g\mathrm{Sh}_{e}] for an edge ee of Γ\Gamma with vertices ss and tt, and hg1Shegh\in g^{-1}\mathrm{Sh}_{e}g. By translating, we may assume g=eg=e, so hSheh\in\mathrm{Sh}_{e}. Corollary 3.5 then implies hh is conjugate to a power of ss or tt. ∎

6. Acylindrical hyperbolicity

In [Vas22], it is shown that (irreducible) 2-dimensional Artin groups of rank at least 3 are acylindrically hyperbolic. By modifying the proof appropriately, we obtain an analogous result for 2-dimensional Shephard groups as an application of Theorem B:

See C

This follows from

Proposition 6.1.

[Vas22, Theorem D] Let XX be a CAT(0)\mathrm{CAT}(0) simplicial complex, together with an action by simplicial isomorphisms of a group GG. Assume that there exists a vertex vv of XX with stabilizer GvG_{v} such that:

  1. (1)

    The orbits of GvG_{v} on the link lk(v,X)lk(v,X) are unbounded, for the associated angular metric.

  2. (2)

    GvG_{v} is weakly malnormal in GG, i.e., there exists an element gGg\in G such that GvgGvg1G_{v}\cap gG_{v}g^{-1} is finite.

Then GG is either virtually cyclic or acylindrically hyperbolic.

Thus the extra assumption (4) of Theorem C is necessary to guarantee condition (1) of Proposition 6.1 is satisfied; otherwise, all links would be finite graphs.

Lemma 6.2.

Let Γ\Gamma be a presentation graph satisfying the hypotheses of Theorem C, and let ee be any edge of Γ\Gamma for which She\mathrm{Sh}_{e} is infinite. Then the orbits of She\mathrm{Sh}_{e} on lk(ve,ΘΓ)lk(v_{e},\Theta_{\Gamma}) are unbounded.

Proof.

Since ΘΓ\Theta_{\Gamma} is 2-dimensional, we know that lk(ve,ΘΓ)Θ^elk(v_{e},\Theta_{\Gamma})\cong\widehat{\Theta}_{e} (see Section 5). Suppose ee has terminal vertices ii and jj, with labels p=pip=p_{i}, m=mijm=m_{ij}, and q=pjq=p_{j}. Since She\mathrm{Sh}_{e} is infinite, we know that 1/p+2/m+1/q11/p+2/m+1/q\leq 1, and She\mathrm{Sh}_{e} is a non-trivial \mathbb{Z}-central extension of a finite index subgroup of a triangle group in either 𝔼2\mathbb{E}^{2} or 2\mathbb{H}^{2} (see Section 3). As discussed in Section 3, the quotient of Θ^e\widehat{\Theta}_{e} by the center of She\mathrm{Sh}_{e} is the 1-skeleton of a semiregular tiling of 𝔼2\mathbb{E}^{2} or 2\mathbb{H}^{2}, and this quotient is a covering map. In particular, the group of deck transformations (which act hyperbolically) of this cover is the central copy of \mathbb{Z} in She\mathrm{Sh}_{e}; thus the orbits of this copy of \mathbb{Z} are unbounded. ∎

In order to show that the Shephard groups in question satisfy (2) of Proposition 6.1, we will detail the portions of the argument which must be modified, and refer the reader to [Vas22, §5] for the full original argument.

Definition 6.3.

Let Γ\Gamma be an extended presentation graph. For vertices s,ts,t of Γ\Gamma, let Γst\Gamma^{st} denote the presentation graph with the same vertex and edge sets and labels as Γ\Gamma, but with the addition of an edge este_{st} labeled 66 if there is no edge between ss and tt in Γ\Gamma. (If there is already an edge between ss and tt, we leave Γ\Gamma unchanged.) We then define the domain KΓstK_{\Gamma^{st}} and metric as in [Vas22, Def. 5.3]. Our complex of groups over KΓstK_{\Gamma^{st}} is defined similarly as well, but we place the free product /ps/pt\mathbb{Z}/p_{s}\mathbb{Z}*\mathbb{Z}/p_{t}\mathbb{Z} as the edge group corresponding to este_{st} if this edge was added; we make no changes if there were no changes made to Γ\Gamma. This is in contrast to the Artin group case, where the rank-2 free group is placed on this edge. We denote the development of this complex of groups by ΘΓst\Theta^{st}_{\Gamma}.

With this modification, the following key lemmas still hold, with completely identical proofs after appropriate replacements are made with the definition(s) above.

Lemma 6.4.

[Vas22, Lemma 5.6] Let ShΓ\mathrm{Sh}_{\Gamma} be a 2-dimensional Shephard group with |V(Γ)|3|V(\Gamma)|\geq 3, and suppose that we are in the second case of [Vas22, Proposition 5.2]. Then ΘΓbc\Theta^{bc}_{\Gamma} is CAT(0)\mathrm{CAT}(0).

Lemma 6.5.

[Vas22, Lemma 5.7] Let ShΓ\mathrm{Sh}_{\Gamma} be a 2-dimensional Shephard group with |V(Γ)|3|V(\Gamma)|\geq 3, such that Γ\Gamma is connected and not right-angled (i.e., has some edge not labeled 22). Then there exists an edge ee of Γ\Gamma between vertices aa and bb with coefficient mab3m_{ab}\geq 3 and an element gSheg\in\mathrm{Sh}_{e} such that ShegSheg1=e\mathrm{Sh}_{e}\cap g\mathrm{Sh}_{e}g^{-1}={e}.

Now we may complete the proof of Theorem C.

Proof (of Theorem C).

We may assume Γ\Gamma is connected, otherwise ShΓ\mathrm{Sh}_{\Gamma} splits as a free product, and each free summand is an infinite group by (4); such groups are always acylindrically hyperbolic. We may assume as well that Γ()\Gamma(\infty) is not right-angled; in this case, ShΓ\mathrm{Sh}_{\Gamma} is a graph product of non-trivial (cyclic) groups, which by [MO15, Cor. 2.13] implies ShΓ\mathrm{Sh}_{\Gamma} is either virtually cyclic or acylindrically hyperbolic. Taking ee to be the edge guaranteed by (4), we know She\mathrm{Sh}_{e} is infinite, not virtually cyclic, and embeds in ShΓ\mathrm{Sh}_{\Gamma}, so under our assumptions ShΓ\mathrm{Sh}_{\Gamma} is not virtually cyclic (hence acylindrically hyperbolic). The proof of [Vas22, Prop. 5.2] and Lemma 6.5 imply that we may choose the edge ee in Lemma 6.5 to be the edge ee guaranteed by (4). We note that this edge has label 3\geq 3, since otherwise She\mathrm{Sh}_{e} would be a direct product of finite groups and hence finite. So Lemmas 6.2 and 6.5 imply ShΓ\mathrm{Sh}_{\Gamma} satisfy the hypotheses of Proposition 6.1, and hence is acylindrically hyperbolic. ∎

7. Relative hyperbolicity and residual finiteness

Recall the following characterization of relative hyperbolicity, due to Bowditch [Bow12]. This phrasing comes from [Osi06, Def. 6.2].

Proposition 7.1.

Let GG be a group and 𝒫={P1,,Pn}\mathcal{P}=\{P_{1},\dots,P_{n}\} a collection of infinite finitely generated subgroups (we will call (G,𝒫)(G,\mathcal{P}) a group pair and the elements of 𝒫\mathcal{P} the peripheral subgroups). Then (G,𝒫)(G,\mathcal{P}) is relatively hyperbolic if and only if GG admits an action on a hyperbolic graph YY such that each of the following hold.

  1. (1)

    All edge stabilizers are finite.

  2. (2)

    All vertex stabilizers are either finite or conjugate to one of the subgroups of 𝒫\mathcal{P}.

  3. (3)

    The number of orbits of edges is finite.

  4. (4)

    The graph YY is fine.

There are a number of equivalent definitions of a fine graph, but the one that shall be useful for us is:

Definition 7.2.

[Bow12, Def 2.1.(F5)] Let YY be a graph and let dYd^{Y} denote the metric on YY which assigns all edges length 11. (By convention, for any graph YY with such a metric, if yy and zz are in different connected components of YY, we say dY(y,z)=d^{Y}(y,z)=\infty.) Fix a vertex xV(Y)x\in V(Y) and let YxY\setminus x denote the largest full subgraph of YY which avoids xx. Let Vx(Y)V_{x}(Y) be the set of vertices of YY which are adjacent to xx. Let dYxd^{Y\setminus x} denote the induced length metric on YxY\setminus x coming from dYd^{Y}, and let dxd_{x} denote the restriction of the metric dYxd^{Y\setminus x} to Vx(Y)V_{x}(Y) (not the induced length metric). We say that YY is fine if the metric space (Vx(Y),dx)(V_{x}(Y),d_{x}) is locally finite222Every finite-radius ball is a finite set. for every xV(Y)x\in V(Y).

In order to show that (certain) Shephard groups ShΓ\mathrm{Sh}_{\Gamma} are relatively hyperbolic, we will use the action of ShΓ\mathrm{Sh}_{\Gamma} on its complex ΘΓ\Theta_{\Gamma}. Specifically, we will let Y=YΓY=Y_{\Gamma} denote the 1-skeleton of ΘΓ\Theta_{\Gamma} endowed with the edge-path metric (each edge is given length 11). When ΘΓ\Theta_{\Gamma} is Gromov hyperbolic, so too is its 1-skeleton as a metric graph under the induced length metric. But this metric graph is quasi-isometric to YY, and hence YY is also hyperbolic. So, our first task is to determine when ΘΓ\Theta_{\Gamma} is hyperbolic.

Lemma 7.3.

Let Γ\Gamma be a 2-dimensional extended presentation graph. If WΓW_{\Gamma} is word hyperbolic, then ΘΓ\Theta_{\Gamma} is Gromov hyperbolic.

Proof.

By Theorem 5.8, ΘΓ\Theta_{\Gamma} is CAT(0)\mathrm{CAT}(0). So by the Flat Plane Theorem, ΘΓ\Theta_{\Gamma} is Gromov hyperbolic if and only if it contains no isometrically embedded copy of 𝔼2\mathbb{E}^{2}. Suppose such an embedded plane exists (so ΘΓ\Theta_{\Gamma} is not hyperbolic). This plane must be a subcomplex of ΘΓ\Theta_{\Gamma}, and in particular must pass through a vertex vv of the form [gSh][g\mathrm{Sh}_{\varnothing}] for some gShΛg\in\mathrm{Sh}_{\Lambda}. This gives rise to an embedded loop of length exactly 2π2\pi in the link of vv. The link of such a vertex is isometric to lk(v,F)Θ^lk(v,K)lk(v_{\varnothing},F_{\varnothing})\ast\widehat{\Theta}_{\varnothing}\cong lk(v_{\varnothing},K) (see discussion after Definition 5.7). As in the Artin group case, such a link contains a circuit of length exactly 2π2\pi if and only if WΓW_{\Gamma} is not hyperbolic [Cri05, Proof of Lemma 5]. In summary, if WΓW_{\Gamma} is hyperbolic, then ΘΓ\Theta_{\Gamma} has no embedded flat plane, and thus ΘΓ\Theta_{\Gamma} is Gromov hyperbolic. ∎

We would like to say that this is an “if and only if” statement, as in the Artin group setting [CC07], but are unsure how to proceed. For the Artin groups, this relies on the existence of an embedding of the Davis complex in the Deligne complex, which is not yet known to exist for arbitrary Shephard groups. However, if such an embedding exists, then the reverse implication is immediate.

Proposition 7.4.

Suppose Γ\Gamma is a 2-dimensional extended presentation graph such that the Davis complex ΣΓ\Sigma_{\Gamma} for the Coxeter group WΓW_{\Gamma} embeds isometrically in the complex ΘΓ\Theta_{\Gamma}. Then if ΘΓ\Theta_{\Gamma} is Gromov hyperbolic, WΓW_{\Gamma} must be word hyperbolic.

Proof.

Suppose WΓW_{\Gamma} is not hyperbolic. Then ΣΓ\Sigma_{\Gamma} is not hyperbolic (since WΓW_{\Gamma} is quasiisometric to ΣΓ\Sigma_{\Gamma}), so it contains an embedded flat plane. Since ΣΓ\Sigma_{\Gamma} embeds isometrically in ΘΓ\Theta_{\Gamma}, we have that ΘΓ\Theta_{\Gamma} also contains an embedded flat plane, so is not hyperbolic. ∎

The existence or non-existence of such an embedding is outside our current scope of consideration. (Although it is natural to conjecture that there is always such an embedding, since it exists for Artin groups and can be constructed case-by-case for finite Shephard groups.)

Now we show that YY is fine. First recall the following notation. If XX is a cell complex and xXx\in X, the open star st(x)=st(x,X)st(x)=st(x,X) is the union of all open cells containing xx, the closed star St(x)=St(x,X)St(x)=St(x,X) is the (topological) closure of the open star, and the boundary of the closed star is St(x)=St(x)st(x)\partial St(x)=St(x)\setminus st(x).

Lemma 7.5.

Suppose Γ\Gamma is 2-dimensional. Let xx be a vertex of ΘΓ\Theta_{\Gamma} (equivalently, of YΓY_{\Gamma}) of the form [gShe][g\mathrm{Sh}_{e}] for gShΓg\in\mathrm{Sh}_{\Gamma} and an edge ee of Γ\Gamma. Let x\ell_{x} denote the induced length metric on St(x)ΘΛ\partial St(x)\subseteq\Theta_{\Lambda}. Then there is a C1C\geq 1 and D0D\geq 0 such that x(y,z)Cdx(y,z)+D\ell_{x}(y,z)\leq Cd_{x}(y,z)+D for all y,zVx(Y)y,z\in V_{x}(Y).

Proof.

Let ι:YΓΘΓ\iota:Y_{\Gamma}\to\Theta_{\Gamma} denote the inclusion map which realizes the quasi-isometry of YΓY_{\Gamma} with the edge length metric and ΘΓ\Theta_{\Gamma} with the Moussong metric. The restriction of ι\iota to YxY\setminus x is a quasi-isometry onto ΘΓst(x)\Theta_{\Gamma}\setminus st(x), under the respective induced length metrics dYxd^{Y\setminus x} and dΘΓst(x)d^{\Theta_{\Gamma}\setminus st(x)}. We choose our CC and DD to be the constants from this restricted quasi-isometry, i.e., those constants which satisfy

1CdYx(y,z)DdΘΓst(x)(y,z)CdYx(y,z)+D\frac{1}{C}d^{Y\setminus x}(y,z)-D\leq d^{\Theta_{\Gamma}\setminus st(x)}(y,z)\leq Cd^{Y\setminus x}(y,z)+D

for all y,zY{x}y,z\in Y\setminus\{x\}. Recall that dxd_{x} is the restriction of dYxd^{Y\setminus x} to Vx(Y)V_{x}(Y), so if we restrict ourselves to elements y,zy,z of Vx(Y)V_{x}(Y), then we can say

1Cdx(y,z)DdΘΓst(x)(y,z)Cdx(y,z)+D\frac{1}{C}d_{x}(y,z)-D\leq d^{\Theta_{\Gamma}\setminus st(x)}(y,z)\leq Cd_{x}(y,z)+D

Fix y,zVx(Y)y,z\in V_{x}(Y). Let γ\gamma be a geodesic in ΘΓst(x)\Theta_{\Gamma}\setminus st(x) from yy to zz. Then

(γ)=dΘΓst(x)(y,z)Cdx(y,z)+D.\ell(\gamma)=d^{\Theta_{\Gamma}\setminus st(x)}(y,z)\leq Cd_{x}(y,z)+D.

Let ρ:ΘΓSt(x)\rho:\Theta_{\Gamma}\to St(x) be the closest point projection onto the closed star St(x)St(x) (the map which sends a point pXp\in X to a point ρ(p)St(x)\rho(p)\in St(x) which minimizes dΘΓ(p,ρ(p))d^{\Theta_{\Gamma}}(p,\rho(p))). Note that St(x)St(x) is a convex set: it is made up of Euclidean right triangles which have one of their acute angles meeting at the common vertex xx. Thus ρ\rho is a well-defined, distance non-increasing retraction [BH13, Prop. II.2.4(4)], and in particular restricts to a well-defined distance non-increasing retraction ρ:ΘΓst(x)St(x)\rho:\Theta_{\Gamma}\setminus st(x)\to\partial St(x). So ρ(γ)\rho(\gamma) is a (rectifiable) path in St(x)\partial St(x) between yy and zz, implying

x(y,z)(ρ(γ))(γ)Cdx(y,z)+D.\ell_{x}(y,z)\leq\ell(\rho(\gamma))\leq\ell(\gamma)\leq Cd_{x}(y,z)+D.\qed

We can now complete the proof of Theorem D.

See D

Proof.

When WΓW_{\Gamma} is hyperbolic, then ΘΓ\Theta_{\Gamma} is hyperbolic by Lemma 7.3. In particular, YY is a hyperbolic graph since it is quasi-isometric to the 1-skeleton of ΘΓ\Theta_{\Gamma}. By Proposition 5.3, the edge stabilizers of ShΓ\mathrm{Sh}_{\Gamma} acting on YY are either finite cyclic (coming from the subgroups generated by the vertices) or trivial. Similarly, the vertex stabilizers are the conjugates of ShΛ\mathrm{Sh}_{\Lambda} for Λ𝒮f\Lambda\in\mathcal{S}^{f}. And since this is a cocompact action with a strict fundamental domain, there are finitely many orbits of edges. Once we show that YY is fine, the result will follow by Proposition 7.1.

Since the action of ShΓ\mathrm{Sh}_{\Gamma} on YY has a strict fundamental domain, we may restrict our consideration to vertices of the form vΛv_{\Lambda} for Λ𝒮f\Lambda\in\mathcal{S}^{f} (using notation from Section 5). Since Γ\Gamma is 2-dimensional, there are three types of vertices x=vΛx=v_{\Lambda} to consider: Λ=\Lambda=\varnothing, Λ={s}\Lambda=\{s\} a single vertex, or Λ=e\Lambda=e a single edge. If Λ=\Lambda=\varnothing, then, as a set, Vx(Y)V_{x}(Y) is simply the vertex set of lk(v,F)lk(v_{\varnothing},F_{\varnothing}), which is finite. Similarly, if Λ={s}\Lambda=\{s\}, then Vx(Y)V_{x}(Y) is the vertex set of the join of lk(v{s},F{s})lk(v_{\{s\}},F_{\{s\}}) and Θ^{s}\widehat{\Theta}_{\{s\}}, which are both finite, and hence Vx(Y)V_{x}(Y) is finite.

Suppose Λ=e\Lambda=e is an edge between vertices ss and tt of Γ\Gamma. Since ΘΓ\Theta_{\Gamma} is a piecewise Euclidean simplicial complex with finitely many isometry types of faces, lk(x,ΘΓ)lk(x,\Theta_{\Gamma}) is isometric to a sufficiently small sphere centered at xx, and this sphere is a radial deformation retract of St(x)\partial St(x). This graph isomorphism is actually a quasi-isometry. We also know that Θ^e\widehat{\Theta}_{e} with the Moussong metric is quasi-isometric to Θ^e\widehat{\Theta}_{e} with the metric dΘ^ed^{\widehat{\Theta}_{e}} which assigns all edges length 11. In particular, there are C1C^{\prime}\geq 1 and D0D^{\prime}\geq 0 such that dΘ^e(y,z)Cx(y,z)+Dd^{\widehat{\Theta}_{e}}(y,z)\leq C^{\prime}\ell_{x}(y,z)+D^{\prime} for all y,zSt(x)y,z\in\partial St(x) (or Θ^e\widehat{\Theta}_{e}).

Now, let CC and DD be the constants guaranteed by Lemma 7.5. Fix a vertex yy of St(x)\partial St(x) (or Vx(Y)V_{x}(Y)). For N>0N>0, define

VN\displaystyle V_{N} ={zVx(Y):dx(y,z)N},\displaystyle=\{\,z\in V_{x}(Y):d_{x}(y,z)\leq N\,\},
LN\displaystyle L_{N} ={zΘ^e:dΘ^e(y,z)C(CN+D)+D}.\displaystyle=\{\,z\in\widehat{\Theta}_{e}:d^{\widehat{\Theta}_{e}}(y,z)\leq C^{\prime}(CN+D)+D^{\prime}\,\}.

Let zVNz\in V_{N}. Then

dΘ^e(y,z)Cx(y,z)+DC(Cdx(y,z)+D)+DC(CN+D)+D.d^{\widehat{\Theta}_{e}}(y,z)\leq C^{\prime}\ell_{x}(y,z)+D^{\prime}\leq C^{\prime}(Cd_{x}(y,z)+D)+D^{\prime}\leq C^{\prime}(CN+D)+D^{\prime}.

x(y,z)Cdx(y,z)+DCN+D\ell_{x}(y,z)\leq Cd_{x}(y,z)+D\leq CN+D, so zSNz\in S_{N}. Hence VNLNV_{N}\subseteq L_{N} for all N>0N>0. We know that Θ^e\widehat{\Theta}_{e} is a locally finite graph: the vertices of Θ^e\widehat{\Theta}_{e} coming from cosets of s\langle s\rangle have valence psp_{s} and the vertices coming from cosets of t\langle t\rangle have valence ptp_{t}, both of which we have assumed to be finite. This implies LNL_{N} is finite for every NN, and hence so is VNV_{N}. Therefore VK(x)V_{K}(x) is locally finite and YΓY_{\Gamma} is fine. ∎

This result is in stark contrast to the Artin group case, where it is very uncommon for Artin groups to be relatively hyperbolic rather than just weakly relatively hyperbolic. We can leverage this not only to show many nice properties of Shephard groups, but also of many Artin groups as well. In the following section, we will discuss an application to Artin groups, but first we will detail some easy consequences of relatively hyperbolicity for Shephard groups.

See E

Proof.

Each property in the list holds for all of ShΓ\mathrm{Sh}_{\Gamma} if and only if it holds for the peripheral subgroups. In more detail:

  1. (1)

    The dihedral Shephard groups are linear, hence have solvable word problem. This implies ShΓ\mathrm{Sh}_{\Gamma} has solvable word problem by [Far98].

  2. (2)

    If (G,𝒫)(G,\mathcal{P}) is a relatively hyperbolic pair, then it was shown in [Tuk94] that any subgroup of GG which does not contain a free group is either finite, virtually infinite cyclic, or is contained in an element of 𝒫\mathcal{P}. Since the peripheral subgroups of ShΛ\mathrm{Sh}_{\Lambda} are linear, they satisfy the Tits alternative, so we can conclude ShΛ\mathrm{Sh}_{\Lambda} does as well.

  3. (3)

    By [BD08], asymptotic dimension of finitely generated groups is preserved by commensurability. Since the 3-dimensional integral Heisenberg group and the universal central extensions of surface groups have finite asymptotic dimension, so too do the peripheral subgroups of ShΛ\mathrm{Sh}_{\Lambda}. By [Osi05], this implies ShΓ\mathrm{Sh}_{\Gamma} has finite asymptotic dimension.

  4. (4)

    If e={i,j}e=\{i,j\} is an edge of Γ\Gamma with 1/pi+2/mij+1/pj=11/p_{i}+2/m_{ij}+1/p_{j}=1, then She\mathrm{Sh}_{e} has polynomial growth since it is commensurable to the 3-dimensional integral Heisenberg group; groups of polynomial growth have the rapid decay property by [Jol90]. If e={i,j}e=\{i,j\} is an edge of Γ\Gamma with 1/pi+2/mij+1/pj<11/p_{i}+2/m_{ij}+1/p_{j}<1, then She\mathrm{Sh}_{e} is a \mathbb{Z}-central extension of a hyperbolic group; such groups have the rapid decay property by [Nos92]. This implies ShΓ\mathrm{Sh}_{\Gamma} has the rapid decay property by [DS05].

Last, if each edge {i,j}\{i,j\} of Γ\Gamma satisfies 1/pi+2/mij+1/pj11/p_{i}+2/m_{ij}+1/p_{j}\not=1, then each peripheral subgroup is biautomatic by Proposition 3.9. This implies ShΛ\mathrm{Sh}_{\Lambda} is biautomatic by [Reb01]. ∎

A more substantial corollary is residual finiteness for certain 2-dimensional Shephard groups and their Artin groups. To discuss residual finiteness, we begin by recalling the notion of a relatively geometric action recently introduced by Einstein and Groves.

Definition 7.6.

[EG22, Def 1.1] Suppose (G,𝒫)(G,\mathcal{P}) is a group pair. An action of GG on a cell complex XX is relatively geometric (with respect to 𝒫)\mathcal{P}) if

  1. (1)

    X/GX/G is compact,

  2. (2)

    Each group in 𝒫\mathcal{P} acts elliptically on XX, and

  3. (3)

    Each stabilizer of a cell in XX is either finite, or conjugate to a finite index subgroup of an element of 𝒫\mathcal{P}.

It is clear from Proposition 5.3 that if G=ShΓG=\mathrm{Sh}_{\Gamma} is a 2-dimensional Shephard group and 𝒫\mathcal{P} is as defined previously, then the action of ShΓ\mathrm{Sh}_{\Gamma} on its complex ΘΓ\Theta_{\Gamma} is relatively geometric with respect to 𝒫\mathcal{P}. We want to make use of:

Proposition 7.7.

[EG22, Cor 1.7] Suppose (G,𝒫)(G,\mathcal{P}) is relatively hyperbolic and acts relatively geometrically (with respect to 𝒫\mathcal{P}) on a CAT(0)\mathrm{CAT}(0) cube complex XX. If every P𝒫P\in\mathcal{P} is residually finite, then GG is residually finite.

We have a characterization of when ShΓ\mathrm{Sh}_{\Gamma} is relatively hyperbolic; if we can determine when ΘΓ\Theta_{\Gamma} is a CAT(0)\mathrm{CAT}(0) cube complex, we will determine a class of residually finite Shephard groups.

Lemma 7.8.

If Γ\Gamma is 2-dimensional and type FC, then ΘΓ\Theta_{\Gamma} is a CAT(0)\mathrm{CAT}(0) cube complex under the “cubical metric”.

Proof.

The cubical metric on the complex ΘΓ\Theta_{\Gamma} is defined as follows: rather than metrize the cell FΛ1FΛ2F_{\Lambda_{1}}\cap F_{\Lambda_{2}}^{*} to be a Coxeter block as in Definition 5.7, we simply give it the metric of a standard Euclidean cube [0,1]n[0,1]^{n} (since this cell is combinatorially a cube). Under this metric, the link of a vertex vΓv_{\Gamma} is still isometric to the spherical join of lk(vΛ,FΛ)lk(v_{\Lambda},F_{\Lambda}) and Θ^Λ\widehat{\Theta}_{\Lambda}, where the metric on these complexes now assigns edge lengths of π/2\pi/2.

In order to show that ΘΓ\Theta_{\Gamma} is a CAT(0)\mathrm{CAT}(0) cube complex, we must show that the links are flag complexes. Since the spherical join of flag complexes is flag, this is equivalent to showing lk(vΛ,FΛ)lk(v_{\Lambda},F_{\Lambda}) and Θ^Λ\widehat{\Theta}_{\Lambda} are flag for all Λ𝒮f\Lambda\in\mathcal{S}^{f}. This is well known for lk(vΛ,FΛ)lk(v_{\Lambda},F_{\Lambda}) since Γ\Gamma is type FC [CD95, Lemma 4.3.4]. If Λ=\Lambda=\varnothing or a single vertex {s}\{s\}, then Θ^Λ\widehat{\Theta}_{\Lambda} is either empty or a finite set, resp., so there is nothing to check. So, suppose Λ\Lambda is an edge between vertices ss and tt. Since Θ^Λ\widehat{\Theta}_{\Lambda} is a graph (i.e., a 1-dimensional simplicial complex), it suffices to show that it contains no 3-cycles. But by Theorem 5.8, the girth of this graph is 2mst4\geq 2m_{st}\geq 4. ∎

Thus we may conclude,

See F

We note that the condition that Γ\Gamma is 2-dimensional and type FC is equivalent to requiring Γ\Gamma has no 3-cycles (sometimes called “triangle-free”). For a triangle-free presentation graph, having no 4-cycles with all edges labeled 22 is equivalent to WΓW_{\Gamma} being hyperbolic (originally due to Moussong in [Mou88], rephrased in terms of the presentation graph in [Cri05, Lemma 5] or [CC07, Prop. 3.1]).

Proof.

Since WΓW_{\Gamma} is hyperbolic, ShΓ\mathrm{Sh}_{\Gamma} is hyperbolic relative to its infinite spherical-type edge subgroups (Theorem D). Since Γ\Gamma is 2-dimensional and type FC, ΘΓ\Theta_{\Gamma} is a CAT(0)\mathrm{CAT}(0) cube complex under the cubical metric (Lemma 7.8) and the action is relatively geometric with respect to the spherical-type edge subgroups. Since the dihedral Shephard groups are residually finite (Proposition 3.7), we conclude that ShΓ\mathrm{Sh}_{\Gamma} is residually finite (Proposition 7.7). ∎

In tandem with Corollary 5.9, we can show

Corollary 7.9.

If Γ\Gamma is triangle-free and has no 4-cycle with all labels 22, then ShΓ\mathrm{Sh}_{\Gamma} is virtually torsion-free.

Proof.

Let sV(Γ)s\in V(\Gamma) be a standard generator of ShΓ\mathrm{Sh}_{\Gamma}. Since ShΓ\mathrm{Sh}_{\Gamma} is residually finite, there is a finite index subgroup Gs<ShΓG_{s}<\mathrm{Sh}_{\Gamma} which avoids ss (meaning sGss\not\in G_{s}), and by taking finitely many intersections, we can choose GsG_{s} to avoid all nontrivial powers of ss. Let G=sV(Γ)GsG=\bigcap_{s\in V(\Gamma)}G_{s}, a finite index subgroup of ShΓ\mathrm{Sh}_{\Gamma} avoiding all nontrivial powers of standard generators. Let N=gShΓg1Gg<GN=\bigcap_{g\in\mathrm{Sh}_{\Gamma}}g^{-1}Gg<G. It is a standard exercise to show that NN (the “normal core” of GG) is finite index when GG is finite index. Suppose gNg\in N has finite order. By Corollary 5.9, we can write g=h1skhg=h^{-1}s^{k}h for some hShΓh\in\mathrm{Sh}_{\Gamma}, some sV(Γ)s\in V(\Gamma), and some kk\in\mathbb{Z}. The definition of NN implies gh1Ghg\in h^{-1}Gh, so skGs^{k}\in G. Since the only power of a generators contained in GG is ee, we must have that sk=es^{k}=e, and consequently g=eg=e. Thus NN is torsion-free. ∎

8. Application to Artin groups

We now establish the analogue of Corollary F for Artin groups. First, we recall the definition of the Deligne complex of an Artin group to establish the notation and terminology which we will use, as it may differ from some references.

Definition 8.1.

Let Γ\Gamma be any presentation graph and AΓ𝒮f={aAΛ:aAΓ,Λ𝒮f}A_{\Gamma}\mathcal{S}^{f}=\{\,aA_{\Lambda}:a\in A_{\Gamma},\Lambda\in\mathcal{S}^{f}\,\} ordered by inclusion. The Deligne complex ΦΓ\Phi_{\Gamma} of AΓA_{\Gamma} is |(AΓ𝒮f)||(A_{\Gamma}\mathcal{S}^{f})^{\prime}|, the geometric realization of the derived complex (AΓ𝒮f)(A_{\Gamma}\mathcal{S}^{f})^{\prime}. An nn-simplex of ΦΓ\Phi_{\Gamma} is written as

[a0AΛ0<a1AΛ1<<anAΛn]\displaystyle[a_{0}A_{\Lambda_{0}}<a_{1}A_{\Lambda_{1}}<\dots<a_{n}A_{\Lambda_{n}}]

where a0AΛ0<a1AΛ1<<anAΛna_{0}A_{\Lambda_{0}}<a_{1}A_{\Lambda_{1}}<\dots<a_{n}A_{\Lambda_{n}} is a chain of elements of AΓ𝒮fA_{\Gamma}\mathcal{S}^{f}. For Λ𝒮Γf\Lambda\in\mathcal{S}^{f}_{\Gamma}, the spherical Deligne complex Φ^Λ\widehat{\Phi}_{\Lambda} is the simplicial complex whose vertices are all cosets of As^A_{\widehat{s}} for sΛs\in\Lambda, with a set of n+1n+1 vertices spanning an nn-simplex if and only if they have nontrivial (global) intersection333By s^\widehat{s}, we mean the largest full subgraph of Λ\Lambda which does not contain ss..

AΓA_{\Gamma} acts on ΦΓ\Phi_{\Gamma} with a strict fundamental domain isomorphic to KΓK_{\Gamma}. Thus we may endow the fundamental domain of ΦΓ\Phi_{\Gamma} with the “same” metric as ΘΓ\Theta_{\Gamma} (the Moussong metric). For the vertex vΛ=[AΛ]v_{\Lambda}=[A_{\Lambda}] of ΦΓ\Phi_{\Gamma}, the link lk(vΛ,ΦΓ)lk(v_{\Lambda},\Phi_{\Gamma}) is isometric to the spherical join

lk(vΛ,FΛ)Φ^Λlk(v_{\Lambda},F_{\Lambda})*\widehat{\Phi}_{\Lambda}

under this metric [CD95], where FΛF_{\Lambda} is defined as in Section 5.

One of the key properties which will allow us to pass between an Artin group and its Shephard quotients is product separability. The following is one of the many equivalent notions of product separability.

Definition 8.2.

Let GG be any group. We say that GG is product separable if for any (finite) collection {G1,,Gn}\{G_{1},\dots,G_{n}\} of finitely generated subgroups of GG, their product G1G2GnG_{1}G_{2}\cdots G_{n} is closed in the profinite topology on GG.

Recall that the profinite topology on GG is the topology whose basis of closed sets are the finite-index subgroups of GG. So it is equivalent to say that G1G2GnG_{1}G_{2}\cdots G_{n} is an intersection of finite-index subgroups of GG.

Lemma 8.3.

Let GG be any group and HH a finite-index subgroup. Then GG is product separable if and only if HH is product separable.

Proof.

Suppose GG is product separable. Let {H1,,Hn}\{H_{1},\dots,H_{n}\} be a collection of finitely generated subgroups of HH. The inclusion map HGH\hookrightarrow G is continuous for any subgroup; the intersection of HH with a finite index subgroup of GG is finite index in HH regardless of the index of HH in GG. In particular, H1H2HnH_{1}H_{2}\cdots H_{n} is closed in GG and is contained in HH, so H1H2HnH_{1}H_{2}\cdots H_{n} is closed in HH.

Now suppose HH is product separable. By replacing HH with the normal core of HH, we may assume HH is normal in GG. (When HH is finite index, so is its normal core, and product separability is inherited by subgroups.) Now let {G1,,Gn}\{G_{1},\dots,G_{n}\} be a collection of finitely generated subgroups of GG, and let Hi=GiHH_{i}=G_{i}\cap H for each ii. Each HiH_{i} is finitely generated since the GiG_{i} are, and HiH_{i} is finite index in GiG_{i} for each ii. So, for each ii, let {h1(i),,hki(i)}\{h_{1}^{(i)},\dots,h_{k_{i}}^{(i)}\} be a system of representatives of the left cosets of HiH_{i} in GiG_{i}. For j=(j1,,jn)j=(j_{1},\dots,j_{n}), let Sj=hj1(1)H1hj2(2)H2hjn(n)HnS_{j}=h_{j_{1}}^{(1)}H_{1}h_{j_{2}}^{(2)}H_{2}\dots h_{j_{n}}^{(n)}H_{n}. Then

G1G2Gn=j=(j1,,jn)Sj.\displaystyle G_{1}G_{2}\cdots G_{n}=\bigcup_{j=(j_{1},\dots,j_{n})}S_{j}.

We claim that each SjS_{j} is closed in GG. For convenience, fix jj and let hi=hji(i)h_{i}=h_{j_{i}}^{(i)}. Then we can rewrite SjS_{j} as

Sj=(h1H1h11)(h1h2H2(h1h2)1)(h1h2h3H3(h1h2h3)1)\displaystyle S_{j}=(h_{1}H_{1}h_{1}^{-1})(h_{1}h_{2}H_{2}(h_{1}h_{2})^{-1})(h_{1}h_{2}h_{3}H_{3}(h_{1}h_{2}h_{3})^{-1})\dots
(h1h2hn1Hn1(h1h2hn1)1)(h1h2hnHn).\displaystyle\dots(h_{1}h_{2}\cdots h_{n-1}H_{n-1}(h_{1}h_{2}\cdots h_{n-1})^{-1})(h_{1}h_{2}\cdots h_{n}H_{n}).

Since HH is normal in GG,

Sj(h1h2hn)1=(h1H1h11)(h1h2H2(h1h2)1)(h1h2h3H3(h1h2h3)1)\displaystyle S_{j}(h_{1}h_{2}\cdots h_{n})^{-1}=(h_{1}H_{1}h_{1}^{-1})(h_{1}h_{2}H_{2}(h_{1}h_{2})^{-1})(h_{1}h_{2}h_{3}H_{3}(h_{1}h_{2}h_{3})^{-1})\dots
(h1h2hnHn(h1h2hn)1)\displaystyle\dots(h_{1}h_{2}\cdots h_{n}H_{n}(h_{1}h_{2}\cdots h_{n})^{-1})

is the product of finitely generated subgroups of HH, thus closed in HH. Since HH is finite index in GG, it is closed in GG as well. The profinite topology is invariant under multiplication (from the left or right), so SjS_{j} is also closed in GG. Thus G1G2GnG_{1}G_{2}\cdots G_{n} is the union of finitely many closed subsets of GG, so is itself closed. ∎

Corollary 8.4.

Dihedral Artin groups are product separable.

Proof.

Suppose AA is a dihedral Artin group with edge label mm. If m=2m=2, then A2A\cong\mathbb{Z}^{2}, and product separability in this case is an easy exercise (following directly from the fact that abelian groups are subgroup separable, i.e., every f.g. subgroup is closed). If m>2m>2, then AA is virtually ×𝔽n\mathbb{Z}\times\mathbb{F}_{n}, where 𝔽n\mathbb{F}_{n} is a rank-nn free group [HJP16, Lemma 4.3]. By [You97], ×𝔽n\mathbb{Z}\times\mathbb{F}_{n} is product separable for any nn, so by Lemma 8.3, so is AA.

By [RZ93, Thm. 2.1], the rank-2 free group 𝔽2\mathbb{F}_{2} is also product separable; this is the dihedral Artin group whose graph is two vertices not joined by an edge. ∎

Lemma 8.5.

If GG is any group and CC is closed in the profinite topology on GG, then for all gCg\not\in C, there exists a finite group FF and a surjective morphism φ:GF\varphi:G\to F such that φ(g)φ(C)\varphi(g)\not\in\varphi(C).

Proof.

Since CC is closed in the profinite topology, it is the intersection of the finite-index subgroups of GG containing it. At least one of these subgroups, say HH, must avoid gg. Let KK denote the normal core of HH, a finite-index normal subgroup of GG contained in HH. Let φ\varphi denote the standard quotient map GG/KG\to G/K. Then gKCKgK\not\subset CK since CKHCK\subseteq H, implying φ(g)φ(C)\varphi(g)\not\in\varphi(C). ∎

The following lemma is the main ingredient in the proof of residual finiteness for these Artin groups.

Lemma 8.6.

Suppose Λ=e\Lambda=e is a single edge with label q<q<\infty. Let x,yΦ^Λx,y\in\widehat{\Phi}_{\Lambda} be any two points. Then there is some N>0N\in\mathbb{Z}_{>0} such that for all k1k\geq 1, the images of xx and yy under the quotient map Φ^ΛΘ^Λ(kN)\widehat{\Phi}_{\Lambda}\to\widehat{\Theta}_{\Lambda(kN)} remain at the same distance.

Proof.

Suppose ss and tt are the vertices of ee, let S=sS=\langle s\rangle and T=tT=\langle t\rangle, and let EE be the edge between SS and TT in Φ^Λ\widehat{\Phi}_{\Lambda}. First, suppose xx and yy are vertices. By symmetry and translating, we may assume x=Sx=S. Let γ=E1E2En\gamma=E_{1}E_{2}\cdots E_{n} be a geodesic edge path connecting xx and yy. Notice that E1E_{1} and EE both contain x=Sx=S; this means there is some k1k_{1}\in\mathbb{Z} such that E1=sk1EE_{1}=s^{k_{1}}E (we could have k1=0k_{1}=0 if E1=EE_{1}=E). Similarly, E2E_{2} and E1E_{1} both contain the vertex sk1Ts^{k_{1}}T, so there is some k2k_{2}\in\mathbb{Z} such that E2=sk1tk2E1E_{2}=s^{k_{1}}t^{k_{2}}E_{1}. Since γ\gamma is a geodesic, it is locally embedded, so k20k_{2}\not=0. Repeating this, for all i>1i>1, we see Ei=αiEi1E_{i}=\alpha_{i}E_{i-1}, where

αi=sk1tk2sk3i syllables,\alpha_{i}=\underbrace{s^{k_{1}}t^{k_{2}}s^{k_{3}}\cdots}_{i\text{ syllables}},

for some ki0k_{i}\not=0. (A syllable is a word of the form sjs^{j} or tjt^{j}.)

Let α=αn\alpha=\alpha_{n}. Notice that

αCSTSn1 cosets,\alpha\not\in C\coloneqq\underbrace{STS\cdots}_{n-1\text{ cosets}},

since otherwise we could reverse the above argument to construct a path from xx to yy with strictly shorter length than γ\gamma. Since AΛA_{\Lambda} is product separable (Corollary 8.4), CC is closed in the profinite topology on AΛA_{\Lambda}, so we can find a finite group FF and surjective morphism φ:AΛF\varphi:A_{\Lambda}\to F so that φ(α)φ(C)\varphi(\alpha)\not\in\varphi(C) (Lemma 8.5). Let psp_{s} and ptp_{t} be the orders of φ(s)\varphi(s) and φ(t)\varphi(t), respectively, and let N=lcm{ps,pt}N=\operatorname{lcm}\{p_{s},p_{t}\}. Fix k1k\geq 1, and let ¯:AΛShΛ(kN)\overline{\,\cdot\,}:A_{\Lambda}\to\mathrm{Sh}_{\Lambda(kN)} denote the standard quotient map. The orders of φ(s)\varphi(s) and φ(t)\varphi(t) divide the orders of s¯\overline{s} and t¯\overline{t}, resp., or in other words, skN,tkNker(φ)s^{kN},t^{kN}\in\mathrm{ker}(\varphi). Since ker(¯)\mathrm{ker}(\overline{\,\cdot\,}) is the normal closure of skNs^{kN} and tkNt^{kN} in AΛA_{\Lambda}, it follows that ker(¯)ker(φ)\mathrm{ker}(\overline{\,\cdot\,})\leq\mathrm{ker}(\varphi). This means there exists a surjection ρ:ShΛ(kN)F\rho:\mathrm{Sh}_{\Lambda(kN)}\to F which makes the diagram

AΛ{{A_{\Lambda}}}F{{F}}ShΛ(kN){{\mathrm{Sh}_{\Lambda(kN)}}}φ\scriptstyle{\varphi}¯\scriptstyle{\overline{\,\cdot\,}}ρ\scriptstyle{\rho}

commute. This implies in particular that α¯C¯\overline{\alpha}\not\in\overline{C}; otherwise φ(α)=ρ(α¯)ρ(C¯)=φ(C)\varphi(\alpha)=\rho(\overline{\alpha})\in\rho(\overline{C})=\varphi(C).

By abuse of notation, let ¯:Φ^ΛΘ^Λ(kN)\overline{\,\cdot\,}:\widehat{\Phi}_{\Lambda}\to\widehat{\Theta}_{\Lambda(kN)} denote the map induced by the quotient AΛShΛ(kN)A_{\Lambda}\to\mathrm{Sh}_{\Lambda(kN)}. We claim that γ¯\overline{\gamma} is still a geodesic between x¯\overline{x} and y¯\overline{y}. Let γ\gamma^{\prime} be a geodesic from x¯\overline{x} to y¯\overline{y} in Θ^Λ(kN)\widehat{\Theta}_{\Lambda(kN)}, say γ=E1Em\gamma^{\prime}=E_{1}^{\prime}\cdots E_{m}^{\prime}. By the same reasoning in Φ^Λ\widehat{\Phi}_{\Lambda}, for each ii we can find i\ell_{i}\in\mathbb{Z} so that Ei=αiEi1E_{i}^{\prime}=\alpha_{i}^{\prime}E_{i-1}^{\prime}, where

αi=s¯1t¯2s¯3i syllables,\alpha_{i}^{\prime}=\underbrace{\overline{s}^{\ell_{1}}\overline{t}^{\ell_{2}}\overline{s}^{\ell_{3}}\cdots}_{i\text{ syllables}},

and i0\ell_{i}\not=0 when i>1i>1. Let α=αm\alpha^{\prime}=\alpha_{m}^{\prime}.

Notice EmE_{m}^{\prime} and E¯n\overline{E}_{n} meet at y¯\overline{y}. Since Θ^Λ(kN)\widehat{\Theta}_{\Lambda(kN)} is bipartite, we know mm and nn have the same parity. If they are odd, let h=th=t and H=TH=T, and if they are even, let h=sh=s and H=SH=S. In either case, y¯\overline{y} is a coset of H¯\overline{H}, so there is some m+1\ell_{m+1}\in\mathbb{Z} such that α¯=αh¯m+1\overline{\alpha}=\alpha^{\prime}\overline{h}^{\ell_{m+1}}. Writing out α\alpha^{\prime}, we see

α¯=(s¯1t¯2s¯3m syllables)h¯m+1=s¯1t¯2s¯3m+1 syllables.\overline{\alpha}=(\underbrace{\overline{s}^{\ell_{1}}\overline{t}^{\ell_{2}}\overline{s}^{\ell_{3}}\cdots}_{m\text{ syllables}})\overline{h}^{\ell_{m+1}}=\underbrace{\overline{s}^{\ell_{1}}\overline{t}^{\ell_{2}}\overline{s}^{\ell_{3}}\cdots}_{m+1\text{ syllables}}.

Since α¯C¯=S¯T¯S¯\overline{\alpha}\not\in\overline{C}=\overline{S}\,\overline{T}\,\overline{S}\cdots (n1n-1 cosets), we know m+1>n1m+1>n-1, i.e., m>n2m>n-2. But since mm and nn are the same parity, mnm\geq n. Thus γ¯\overline{\gamma} is a geodesic and d(x¯,y¯)=(γ¯)=(π/q)n=(γ)=d(x,y)d(\overline{x},\overline{y})=\ell(\overline{\gamma})=(\pi/q)n=\ell(\gamma)=d(x,y).

Now suppose xx and yy are any two points of Φ^Λ\widehat{\Phi}_{\Lambda}. If xx is a vertex, let X1=X2=xX_{1}=X_{2}=x, and if xx is not a vertex, let X1X_{1} and X2X_{2} be the two distinct vertices of the edge containing xx. Similarly, if yy is a vertex, let Y1=Y2=yY_{1}=Y_{2}=y, and if yy is not a vertex, let Y1Y_{1} and Y2Y_{2} be the two distinct vertices of the edge containing yy. For i,j{1,2}i,j\in\{1,2\}, let γij\gamma_{ij} be a geodesic from XiX_{i} to YjY_{j}. For each i,ji,j, by our work above we can find Nij1N_{ij}\geq 1 so that γij\gamma_{ij} remains a geodesic of the same length under the quotient to Θ^Λ(kNij)\widehat{\Theta}_{\Lambda(kN_{ij})} for each k1k\geq 1. Let N=lcm{Nij}N=\operatorname{lcm}\{N_{ij}\} and fix k1k\geq 1. Let ¯:Φ^ΛΘ^Λ(kN)\overline{\,\cdot\,}:\widehat{\Phi}_{\Lambda}\to\widehat{\Theta}_{\Lambda(kN)} be the quotient map. Then each γij¯\overline{\gamma_{ij}} is a geodesic in Θ^Λ(kN)\widehat{\Theta}_{\Lambda(kN)} of the same length as γij\gamma_{ij}. Suppose γ\gamma is a geodesic from x¯\overline{x} to y¯\overline{y} in Θ^Λ(kN)\widehat{\Theta}_{\Lambda(kN)}. Then there is some II and some JJ so that γ\gamma passes through XI¯\overline{X_{I}} and YJ¯\overline{Y_{J}}. Note that since ¯\overline{\,\cdot\,} is a simplicial map, d(x¯,XI¯)=d(x,XI)d(\overline{x},\overline{X_{I}})=d(x,X_{I}) and d(y¯,YJ¯)=d(y,YJ)d(\overline{y},\overline{Y_{J}})=d(y,Y_{J}). Then

d(x¯,y¯)\displaystyle d(\overline{x},\overline{y}) =(γ)\displaystyle=\ell(\gamma)
=(γ|[x¯,X¯I])+(γ|[X¯I,Y¯J])+(γ|[Y¯J,y¯])\displaystyle=\ell(\gamma|_{[\overline{x},\overline{X}_{I}]})+\ell(\gamma|_{[\overline{X}_{I},\overline{Y}_{J}]})+\ell(\gamma|_{[\overline{Y}_{J},\overline{y}]})
=d(x¯,X¯I)+(γ|[X¯I,Y¯J])+d(Y¯J,y¯)\displaystyle=d(\overline{x},\overline{X}_{I})+\ell(\gamma|_{[\overline{X}_{I},\overline{Y}_{J}]})+d(\overline{Y}_{J},\overline{y})
=d(x¯,X¯I)+(γ¯IJ)+d(Y¯J,y¯)\displaystyle=d(\overline{x},\overline{X}_{I})+\ell(\overline{\gamma}_{IJ})+d(\overline{Y}_{J},\overline{y})
=d(x,XI)+(γIJ)+d(YJ,y)\displaystyle=d(x,X_{I})+\ell(\gamma_{IJ})+d(Y_{J},y)
=d(x,XI)+d(XI,YJ)+d(YJ,y)\displaystyle=d(x,X_{I})+d(X_{I},Y_{J})+d(Y_{J},y)
d(x,y).\displaystyle\geq d(x,y).

Since the quotient clearly cannot increase distance, d(x,y)=d(x¯,y¯)d(x,y)=d(\overline{x},\overline{y}). ∎

Lemma 8.7.

Suppose Γ\Gamma is a 2-dimensional presentation graph. Let γ\gamma be a geodesic segment of ΦΓ\Phi_{\Gamma}. Then there is some k2k\geq 2 so that the image of γ\gamma under the natural quotient ΦΓΘΓ(k)\Phi_{\Gamma}\to\Theta_{\Gamma(k)} remains a geodesic segment of the same length.

Proof.

For k2k\geq 2, let ρk:ΦΓΘΓ(k)\rho_{k}:\Phi_{\Gamma}\to\Theta_{\Gamma(k)} denote the usual quotient map induced by the quotient qk:AΓShΓ(k)q_{k}:A_{\Gamma}\to\mathrm{Sh}_{\Gamma(k)}. Let xx be a point in the interior of γ\gamma.

If xx is in the interior of a 2-simplex of ΦΓ\Phi_{\Gamma}, then ρk(γ)\rho_{k}(\gamma) is locally geodesic at ρk(x)\rho_{k}(x) for any value of kk, since ρp\rho_{p} maps 2-simplices isometrically to 2-simplices. In this case, define nx=2n_{x}=2.

If there is a neighborhood of xx in γ\gamma which is contained in an edge, then there is nothing to show since edges are mapped isometrically to edges (in other words, ρk(γ)\rho_{k}(\gamma) is locally geodesic at ρk(x)\rho_{k}(x) for any value of kk). In this case, as above, define nx=2n_{x}=2.

Suppose xx is a point contained in the interior of an edge EE of ΦΓ\Phi_{\Gamma} with no neighborhood of xx in γ\gamma also contained in this edge. First, suppose E=[a1A,a2As]E=[a_{1}A_{\varnothing},a_{2}A_{s}] for a vertex ss of Γ\Gamma or E=[a1A,a2Ae]E=[a_{1}A_{\varnothing},a_{2}A_{e}] for an edge ee of Γ\Gamma. Then the stabilizer of EE is trivial, and in particular, the link of EE is mapped isomorphically to the link of ρk(E)\rho_{k}(E) for any kk. So suppose E=[a1As,a2Ae]E=[a_{1}A_{s},a_{2}A_{e}] for a vertex ss of Γ\Gamma and an edge ee of Γ\Gamma containing ss. Let Δ1\Delta_{1} and Δ2\Delta_{2} denote the distinct 2-simplices containing EE whose interiors nontrivially intersect Γ\Gamma. The stabilizer of EE is conjugate to s\langle s\rangle and acts transitively on the link of EE, so there is some gAΓg\in A_{\Gamma} and some nxn_{x}\in\mathbb{Z} such that Δ2=(g1snxg)Δ1\Delta_{2}=(g^{-1}s^{n_{x}}g)\Delta_{1}. (Since γ\gamma is a geodesic, nx0n_{x}\not=0.) For k>|nx|k>|n_{x}|, ρk(Δ1)\rho_{k}(\Delta_{1}) and ρk(Δ2)\rho_{k}(\Delta_{2}) are the 2-simplices of ΘΓ(k)\Theta_{\Gamma(k)} containing ρk(E)\rho_{k}(E) (and ρk(x)\rho_{k}(x)) whose interiors intersect ρk(γ)\rho_{k}(\gamma). Moreover, ρk(Δ2)=qk(g1snxg)ρk(Δ1)\rho_{k}(\Delta_{2})=q_{k}(g^{-1}s^{n_{x}}g)\rho_{k}(\Delta_{1}). If ρk(Δ2)=ρk(Δ1)\rho_{k}(\Delta_{2})=\rho_{k}(\Delta_{1}), then qk(g1snxg)=eq_{k}(g^{-1}s^{n_{x}}g)=e, but this happens if and only if qk(snx)=eq_{k}(s^{n_{x}})=e. Since k>|nx|k>|n_{x}|, this can’t happen, so ρk(Δ2)ρk(Δ1)\rho_{k}(\Delta_{2})\not=\rho_{k}(\Delta_{1}) and ρk(γ)\rho_{k}(\gamma) is locally geodesic at ρk(x)\rho_{k}(x) for all k>|nx|k>|n_{x}|.

Last, suppose xx is a vertex of ΦΓ\Phi_{\Gamma}. Then x=[gAΛ]x=[gA_{\Lambda}] for some gAΓg\in A_{\Gamma} and some Λ𝒮f\Lambda\in\mathcal{S}^{f}. There are three subcases to consider.

First, assume Λ=\Lambda=\varnothing. Then for any k2k\geq 2, ρk(x)=[qk(g)Sh]\rho_{k}(x)=[q_{k}(g)\mathrm{Sh}_{\varnothing}], and ρk\rho_{k} induces an isometry of the link of xx and the link of ρk(x)\rho_{k}(x) (both being isometric to lk(v,F)lk(v_{\varnothing},F_{\varnothing})).

Next, assume Λ={s}\Lambda=\{s\} for a vertex ss of Γ\Gamma. The link of xx is the join of lk(vs,Fs)lk(v_{s},F_{s}) and Φ^{s}\widehat{\Phi}_{\{s\}}, the latter of which is in bijection with s\langle s\rangle\cong\mathbb{Z}. The intersection of the ε\varepsilon-sphere of xx with γ\gamma induces two points y,zlk(x)y,z\in lk(x) of distance π\geq\pi apart. Every point of Φ^{s}\widehat{\Phi}_{\{s\}} is distance π/2\pi/2 from every point of lk(vs,Fs)lk(v_{s},F_{s}) (by the definition of spherical join), so either y,zlk(vs,Fs)y,z\in lk(v_{s},F_{s}) or y,zΦ^{s}y,z\in\widehat{\Phi}_{\{s\}}. If y,zlk(vs,Fs)y,z\in lk(v_{s},F_{s}), then as in the previous case ρk\rho_{k} acts isometrically on lk(vs,Fs)lk(v_{s},F_{s}) for any k2k\geq 2, so yy and zz remain at the same distance under ρk\rho_{k}. If y,zΦ^{s}y,z\in\widehat{\Phi}_{\{s\}}, then they can be represented by some powers of ss, say sns^{n} and sms^{m} with nmn\not=m\in\mathbb{Z}. (Since γ\gamma is a geodesic, these points are distinct.) By translating we may assume these points are in fact ee (=s0=s^{0}) and snxs^{n_{x}} for nx=mn0n_{x}=m-n\not=0. Choosing k>|nx|k>|n_{x}|, ρk\rho_{k} acts by the standard quotient /k\mathbb{Z}\to\mathbb{Z}/k\mathbb{Z} and thus these points remain distinct. In other words, for k>|nx|k>|n_{x}|, ρk(γ)\rho_{k}(\gamma) is locally geodesic at ρk(x)\rho_{k}(x).

Finally, suppose Λ\Lambda is an edge with finite label mm. Then, as before, the intersection of γ\gamma with the ε\varepsilon-sphere of xx induces two points y,zlk(x)=Φ^Λy,z\in lk(x)=\widehat{\Phi}_{\Lambda} of distance π\geq\pi. Lemma 8.6 implies that there is some nx1n_{x}\geq 1 so that the distance between yy and zz remains π\geq\pi in the quotient via ρk\rho_{k} for all kk which are (positive) multiples of nxn_{x}. In other words, for all such kk, ρk(γ)\rho_{k}(\gamma) is locally geodesic at ρk(x)\rho_{k}(x).

Let k2k\geq 2 be a common multiple of {|nx|:xint(γ)}\{\,|n_{x}|:x\in\mathrm{int}(\gamma)\,\}. (This set is finite and consists of positive integers, so kk exists.) Our above arguments show that ρk(γ)\rho_{k}(\gamma) is locally geodesic at ρk(x)\rho_{k}(x) for all xx, or in other words, that ρk(γ)\rho_{k}(\gamma) is a local geodesic in ΘΓ(k)\Theta_{\Gamma(k)}. Since ΘΓ(k)\Theta_{\Gamma(k)} is CAT(0)\mathrm{CAT}(0), this implies ρk(γ)\rho_{k}(\gamma) is a geodesic and has the same length as γ\gamma. ∎

Now we may complete the proof of

See G

Proof.

Let gAΓ{e}g\in A_{\Gamma}\setminus\{e\}. Let γ\gamma be the geodesic in ΦΓ\Phi_{\Gamma} from [A][A_{\varnothing}] to [gA][gA_{\varnothing}]. By Lemma 8.7, there is some kk so that the image γ¯\overline{\gamma} of γ\gamma under the quotient map to ΘΓ(k)\Theta_{\Gamma(k)} is a geodesic. If g¯\overline{g} denotes the image of gg under the quotient AΓShΓ(k)A_{\Gamma}\to\mathrm{Sh}_{\Gamma(k)}, then γ¯\overline{\gamma} is a geodesic from [Sh][\mathrm{Sh}_{\varnothing}] to [g¯Sh][\overline{g}\mathrm{Sh}_{\varnothing}]. Since ΘΓ(k)\Theta_{\Gamma(k)} is CAT(0)\mathrm{CAT}(0), this means g¯e\overline{g}\not=e. But ShΓ(k)\mathrm{Sh}_{\Gamma(k)} is residually finite (Corollary F), so there is a further quotient ShΓ(k)F\mathrm{Sh}_{\Gamma(k)}\to F to a finite group FF under which g¯\overline{g} remains nontrivial. Composing these maps gives a quotient AΓFA_{\Gamma}\to F to a finite group where the image of gg is nontrivial. ∎

References

  • [AS83] Kenneth I Appel and Paul E Schupp. Artin groups and infinite Coxeter groups. Inventiones mathematicae, 72(2):201–220, 1983.
  • [BD08] Greg Bell and Alexander Dranishnikov. Asymptotic dimension. Topology and its Applications, 155(12):1265–1296, 2008.
  • [BH13] Martin R Bridson and André Haefliger. Metric spaces of non-positive curvature, volume 319. Springer Science & Business Media, 2013.
  • [BM10] Tom Brady and Jon McCammond. Braids, posets and orthoschemes. Algebraic & Geometric Topology, 10(4):2277–2314, 2010.
  • [Bow12] Brian H Bowditch. Relatively hyperbolic groups. International Journal of Algebra and Computation, 22(03):1250016, 2012.
  • [Bro82] Kenneth S Brown. Cohomology of Groups. Springer Science & Business Media, 1982.
  • [CC07] Ruth Charney and John Crisp. Relative hyperbolicity and Artin groups. Geometriae Dedicata, 129:1–13, 2007.
  • [CD95] Ruth Charney and Michael W Davis. The K(π,1)K(\pi,1)-problem for hyperplane complements associated to infinite reflection groups. Journal of the American Mathematical Society, 8(3):597–627, 1995.
  • [Cox75] H.S.M. Coxeter. Regular Complex Polytopes. Cambridge University Press, 1975.
  • [Cri05] John Crisp. Automorphisms and abstract commensurators of 2–dimensional Artin groups. Geometry & Topology, 9(3):1381–1441, 2005.
  • [DK18] Cornelia Druţu and Michael Kapovich. Geometric group theory, volume 63. American Mathematical Soc., 2018.
  • [dLH00] Pierre de La Harpe. Topics in geometric group theory. University of Chicago Press, 2000.
  • [DS05] Cornelia Druţu and Mark Sapir. Relatively hyperbolic groups with rapid decay property. International Mathematics Research Notices, 2005(19):1181–1194, 2005.
  • [EG22] Eduard Einstein and Daniel Groves. Relatively geometric actions on CAT(0)\mathrm{{CAT}}(0) cube complexes. Journal of the London Mathematical Society, 105(1):691–708, 2022.
  • [Far98] Benson Farb. Relatively hyperbolic groups. Geometric and functional analysis, 8(5):810–840, 1998.
  • [Gol23] Katherine Goldman. CAT(0) and cubulated Shephard groups. arXiv preprint arXiv:2310.10883, 2023.
  • [Hae22] Thomas Haettel. XXL type Artin groups are CAT(0) and acylindrically hyperbolic. In Annales de l’Institut Fourier, volume 72, pages 2541–2555, 2022.
  • [Hat02] Allen Hatcher. Algebraic Topology. Cambridge University Press, 2002.
  • [HJP16] Jingyin Huang, Kasia Jankiewicz, and Piotr Przytycki. Cocompactly cubulated 2-dimensional Artin groups. Commentarii Mathematici Helvetici, 91(3):519–542, 2016.
  • [Jan22] Kasia Jankiewicz. Residual finiteness of certain 2-dimensional Artin groups. Advances in Mathematics, 405:108487, 2022.
  • [Jol90] Paul Jolissaint. Rapidly decreasing functions in reduced C{C}^{*}-algebras of groups. Transactions of the American Mathematical Society, 317(1):167–196, 1990.
  • [Kal] Generated with KaleidoTile by Jeff Weeks:
    https://www.geometrygames.org/KaleidoTile/index.html.en.
  • [KS04] Ilya Kapovich and Paul Schupp. Relative hyperbolicity and Artin groups. Geometriae Dedicata, 107:153–167, 2004.
  • [Mil75] John Milnor. On the 3-dimensional Brieskorn manifolds M(p,q,r){M}(p,q,r). Knots, groups and 3-Manifolds, 3:175–225, 1975.
  • [MO15] Ashot Minasyan and Denis Osin. Acylindrical hyperbolicity of groups acting on trees. Mathematische Annalen, 362(3):1055–1105, 2015.
  • [Mou88] Gabor Moussong. Hyperbolic Coxeter groups. PhD thesis, The Ohio State University, 1988.
  • [MS16] Giovanni Moreno and Monika Ewa Stypa. On the vertex-to-edge duality between the Cayley graph and the coset geometry of von Dyck groups. Mathematica Slovaca, 66(3):527–538, 2016.
  • [Nos92] Gennady Andreevich Noskov. Algebras of rapidly decreasing functions on groups and cocycles of polynomial growth. Sibirskii Matematicheskii Zhurnal, 33(4):97–103, 1992.
  • [NR97] Walter D Neumann and Lawrence Reeves. Central extensions of word hyperbolic groups. Annals of mathematics, 145(1):183–192, 1997.
  • [Osi05] Denis Osin. Asymptotic dimension of relatively hyperbolic groups. International Mathematics Research Notices, 2005(35):2143–2161, 2005.
  • [Osi06] Denis Osin. Relatively Hyperbolic Groups: Intrinsic Geometry, Algebraic Properties, and Algorithmic Problems, volume 843. American Mathematical Soc., 2006.
  • [Reb01] Donovan Yves Rebbechi. Algorithmic properties of relatively hyperbolic groups. PhD thesis, 2001.
  • [RZ93] Luis Ribes and Pavel A Zalesskii. On the profinite topology on a free group. Bulletin of the London Mathematical Society, 25(1):37–43, 1993.
  • [She52] G. C. Shephard. Regular complex polytopes. Proceedings of the London Mathematical Society, s3-2(1):82–97, 1952.
  • [Tuk94] Pekka Tukia. Convergence groups and Gromov’s metric hyperbolic spaces. New Zealand J. Math., 23(2):157, 1994.
  • [Vas22] Nicolas Vaskou. Acylindrical hyperbolicity for Artin groups of dimension 2. Geometriae Dedicata, 216(1):7, 2022.
  • [You97] Shihong You. The product separability of the generalized free product of cyclic groups. Journal of the London Mathematical Society, 56(1):91–103, 1997.