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

Virtually cocompactly cubulated Artin-Tits groups

Thomas Haettel

Abstract. We give a conjectural classification of virtually cocompactly cubulated Artin-Tits groups (i.e. having a finite index subgroup acting geometrically on a CAT(0) cube complex), which we prove for all Artin-Tits groups of spherical type, FC type or two-dimensional type. A particular case is that for n4n\geqslant 4, the nn-strand braid group is not virtually cocompactly cubulated.

footnotetext: Keywords : Artin-Tits groups, braid groups, CAT(0) cube complexes. AMS codes : 20F36, 20F65, 20F67

Introduction

Groups acting geometrically on CAT(0) spaces (called CAT(0) groups), or even better on CAT(0) cube complexes (called cocompactly cubulated groups), possibly up to a finite index subgroup, enjoy a list of nice properties: they have a quadratic Dehn function, a solvable word and conjugacy problem, they have the Haagerup property, their amenable subgroups are virtually abelian and undistorted, they satisfy the Tits alternative… R. Charney conjectures that all Artin-Tits groups are CAT(0), but very few cases are known. With D. Kielak and P. Schwer (see [HKS16]), we pursued the construction of T. Brady and J. McCammond (see [Bra01] and [BM10]) to prove that for n6n\leqslant 6, the nn-strand braid group is CAT(0).

In this article, we give a conjectural classification of which Artin-Tits groups are virtually cocompactly cubulated, and we prove this classification under a mild conjecture on Artin-Tits groups, which is satisfied in particular for spherical, type FC or 22-dimensional Artin-Tits groups. Right-angled Artin groups are well-known to act cocompactly on their Salvetti CAT(0) cube complex, but there are a few more examples. This question was asked by D. Wise for the particular case of braid groups (see [Wis, Problem 13.4]).

Conjecture A (Classification of virtually cocompactly cubulated Artin-Tits groups).

Let M=(mab)a,bSM=(m_{ab})_{a,b\in S} be a finite Coxeter matrix. Then the Artin-Tits group A(M)A(M) is virtually cocompactly cubulated if and only if the following two conditions are satisfied:

  1. 1.

    for each pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} is odd, either mac=mbc=m_{ac}=m_{bc}=\infty or mac=mbc=2m_{ac}=m_{bc}=2, and

  2. 2.

    for each distinct a,bSa,b\in S such that mabm_{ab} is even and different from 22, there is an ordering of {a,b}\{a,b\} (say a<ba<b) such that, for every cS{a,b}c\in S\char 92\relax\{a,b\}, one of the following holds:

    • \bullet

      mac=mbc=2m_{ac}=m_{bc}=2,

    • \bullet

      mac=2m_{ac}=2 and mbc=m_{bc}=\infty,

    • \bullet

      mac=mbc=m_{ac}=m_{bc}=\infty, or

    • \bullet

      macm_{ac} is even and different from 22, a<ca<c in the ordering of {a,c}\{a,c\}, and mbc=m_{bc}=\infty.

In particular, typical examples of cocompactly cubulated Artin-Tits groups are the following types:

  • \bullet

    right-angled Artin groups, i.e. such that a,bS,mab{2,}\forall a,b\in S,m_{ab}\in\{2,\infty\},

  • \bullet

    dihedral Artin groups, i.e. such that |S|=2|S|=2,

  • \bullet

    “even stars” Artin groups, i.e. such that there exists a “central vertex” a0Sa_{0}\in S such that a,bS{a0},mab=\forall a,b\in S\char 92\relax\{a_{0}\},m_{ab}=\infty and aS{a0},maa0\forall a\in S\char 92\relax\{a_{0}\},m_{aa_{0}} is even.

You can see in Figure 1 an example of the Coxeter graph of an even star Artin-Tits group. In that figure, all the edges labeled \infty are not drawn.

Refer to captiona0a_{0}44Refer to caption446666221010Refer to caption
Figure 1: An example of the Coxeter graph of an “even star” cocompactly cubulated Artin-Tits group, with central vertex a0a_{0}

The most general picture of an arbitrary cocompactly cubulated Artin-Tits group comes roughly from combining dihedral Artin groups and even stars Artin groups, in a right-angled-like fashion.

Another way to state Conjecture A is by describing local obstructions in the Coxeter matrix MM, see also Figure 2. In particular, standard parabolic subgroups of rank 33 and 44 should determine if an Artin-Tits group is virtually cocompactly cubulated or not. Recall that the rank of an Artin group (or a standard parabolic subgroup) is the number of standard generators.

Conjecture B (Reformulation of Conjecture A).

Let M=(mab)a,bSM=(m_{ab})_{a,b\in S} be a finite Coxeter matrix. Then the Artin-Tits group A(M)A(M) is not virtually cocompactly cubulated if and only one of the following occurs:

  • \bullet

    there exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} is odd, macm_{ac}\neq\infty and mbc2m_{bc}\neq 2,

  • \bullet

    there exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} and macm_{ac} are even numbers different from 22, and mbcm_{bc}\neq\infty, or

  • \bullet

    there exist 44 pairwise distinct a,b,c,dSa,b,c,d\in S such that mab{2,}m_{ab}\not\in\{2,\infty\}, mac,mbdm_{ac},m_{bd}\neq\infty and mad,mbc2m_{ad},m_{bc}\neq 2.

Refer to captionoddRefer to captionbbccaa2\neq 2\neq\inftyRefer to captioneven 2\neq 2Refer to captionbbccaaeven 2\neq 2\neq\inftyRefer to caption2,\neq 2,\inftyRefer to captionaabbcc2\neq 2\neq\inftyRefer to caption\neq\infty2\neq 2dd
Figure 2: Local obstructions in Conjecture B

One implication of Conjecture A, namely the cubulation of Artin-Tits groups satisfying the two conditions, is proven in full generality in this article.

Theorem C.

Let M=(mab)a,bSM=(m_{ab})_{a,b\in S} be a finite Coxeter matrix satisfying the two conditions of Conjecture A. Then A(M)A(M) is cocompactly cubulated.

The converse implication of Conjecture A, namely to show that the two conditions are necessary to be virtually cocompactly cubulated, is proven under the following mild assumption. Let M=(mab)a,bSM=(m_{ab})_{a,b\in S} be a finite Coxeter matrix. We say that the Artin-Tits group A(M)A(M) satisfies property ()(\dagger) if

sS,n1,ZA(M)(sn)=ZA(M)(s).\forall s\in S,\forall n\geqslant 1,Z_{A(M)}(s^{n})=Z_{A(M)}(s).
Theorem D.

Let M=(mab)a,bSM=(m_{ab})_{a,b\in S} be a finite Coxeter matrix such that A(M)A(M) satisfies property ()(\dagger). If A(M)A(M) is virtually cocompactly cubulated, then A(M)A(M) satisfies the two conditions of Conjecture A.

It is conjectured that all Artin-Tits groups satisfy property ()(\dagger), it is notably a very restricted consequence of Property ()(\star) in [God07]. In particular, it is true as soon as the Deligne complex can be endowed with a piecewise Euclidean CAT(0) metric. This condition is therefore true for Artin-Tits groups of type FC (i.e. every complete subgraph spans a spherical subgroup), with the cubical metric on the Deligne complex (see [CD95b, Theorem A]). It is also true if the Artin-Tits group A(M)A(M) is such that any irreducible spherical parabolic subgroup has rank at most 22 (which is slightly more general than 22-dimensional), in which case the Moussong metric on the Deligne complex is CAT(0) (see [CD95b, Theorem A]). Note that the Moussong metric on the Deligne complex is conjectured to be CAT(0) for all Artin-Tits groups (see [CD95b, Conjecture 3]).

Theorem E.

Conjecture A holds for any Artin-Tits group satisfying property ()(\dagger). In particular, Conjecture A holds for Artin-Tits groups of type FC, and for Artin-Tits groups whose irreducible spherical parabolic subgroups have rank at most 22.

One should note that the two conditions in Conjecture A imply that the Artin-Tits group is of type FC. In particular, the following is a consequence of Conjecture A.

Conjecture F (Consequence of Conjecture A).

If an Artin-Tits group is virtually cocompactly cubulated, then it is of type FC.

In the particular case or Artin groups of spherical type, the condition is much simpler.

Corollary G (Classification of virtually cocompactly cubulated Artin-Tits groups of spherical type).

Let AA be an Artin-Tits group of spherical type. Then AA is virtually cocompactly cubulated if and only if every irreducible parabolic subgroup of AA has rank at most 22.

In particular, this gives a very simple answer for braid groups.

Corollary H (Cubulation of braid groups).

The nn-strand braid group BnB_{n}, or its central quotient Bn/Z(Bn)B_{n}/Z(B_{n}), is virtually cocompactly cubulated if and only if n3n\leqslant 3.

However, according to B. Bowditch (see [Bow13]), all mapping class groups, including braid groups, are coarse median, which implies that their asymptotic cones “look like ” asymptotic cones of CAT(0) cube complexes: they are not cocompactly cubulated, but “look cubical ” on a large scale.

Concerning proper actions of Artin groups on CAT(0) cube complexes, even the following question is still open.

Question (Charney [Cha], Wise [Wis]).

Does the 44-strand braid group B4B_{4} have a metrically proper action on a CAT(0) cube complex ?

During the proof, we also prove the following cubulation results, of independent interest. See Theorem 2.1 and Proposition 4.2 fore more precise versions.

Theorem I (Cubulation of normalizers and centralizers).

Let GG be a cocompactly cubulated group, and let AA be an abelian subgroup of GG. Then AA has a finite index subgroup A0A_{0} such that NG(A0)N_{G}(A_{0}) is cocompactly cubulated, and ZG(A0)Z_{G}(A_{0}) has finite index in NG(A0)N_{G}(A_{0}).

Theorem J (Cubulation of central quotients).

Let GG be a cocompactly cubulated group, and let AA be a central, convex-cocompact subgroup of GG. Then G/AG/A is cocompactly cubulated.

In an earlier version of this article, we obtained only Theorem D for cocompactly cubulated Artin groups, without the virtual part. J. Huang, K. Jankiewicz and P. Przytycki, simultaneously to this earlier version and independently, proved Theorem D for 22-dimensional Artin groups with the virtual part (see [HJP16]). In particular, they showed that a 22-dimensional Artin group is cocompactly cubulated if and only if it is virtually cocompactly cubulated.

Concerning Coxeter groups, Niblo and Reeves proved (see [NR03]) that every Coxeter group acts properly on a locally finite CAT(0) cube complex. Caprace and Mühlherr proved (see [CM05]) that this action is cocompact if and only if the Coxeter diagram does not contain an affine subdiagram of rank at least 33.

O. Varghese recently described (see [Var15]) a group-theoretic condition ensuring that any (strongly simplicial) isometric action on a CAT(0) cube complex has a global fixed point. This condition is notably satisfied by Aut(𝔽n)Aut(\mathbb{F}_{n}) for n1n\geqslant 1.

Outline of the proof The rough idea is to study the CAT(0) visual angle between maximal abelian subgroups in Artin groups. Using a result from J. Crisp and L. Paoluzzi (see [CP05]), we show that if a,ba,b are the standard generators of the 33-strand braid group acting on some CAT(0) space, then the translation axes for aa and abababababab form an acute visual angle at infinity.

On the other hand, we show that the translation axes of elements in maximal abelian subgroups of a group acting geometrically on a CAT(0) cube complex, with finite intersection, form an obtuse visual angle at infinity. This is the source of the non-cubicality results. This uses a flat torus theorem for maximal abelian subgroups of cocompactly cubulated groups by Wise and Woodhouse (see [WW17] and Theorem 3.6).

Acknowledgments: The author would like to thank very warmly Jingyin Huang, Kasia Jankiewicz and Piotr Przytycki for their precious help to find and correct several mistakes in previous versions of this article. The author would also like to thank Daniel Wise for inspiring discussions. The author would also like to thank Eddy Godelle and Luis Paris, for very interesting discussions on Artin groups. The author would also like to thank Nir Lazarovich and Anthony Genevois for noticing several typos and a few mistakes. Finally, the author would like to thank several anonymous referees for pointing out mistakes in previous versions of this article, and for many useful comments that helped in particular to improve the exposition.

1 Definitions and notations

1.1 Artin groups

For pp\in\mathbb{N}, let wpw_{p} denote the word wp(a,b)=ababaw_{p}(a,b)=aba...ba of length pp. Let SS be a finite set, and let Γ\Gamma be a graph with vertex set SS and edges labeled in 2\mathbb{N}_{\geqslant 2}. The Artin-Tits group A(Γ)A(\Gamma) is defined by the following presentation:

A(Γ)=sS|wp(s,t)=wp(t,s) for each edge {s,t} labeled p.A(\Gamma)=\left\langle s\in S\,|\,w_{p}(s,t)=w_{p}(t,s)\mbox{ for each edge $\{s,t\}$ labeled $p$}\right\rangle.

If S={a,b}S=\{a,b\}, then A(Γ)A(\Gamma) is called a dihedral Artin group, and we will denote it by A(p)A(p), where pp is the label of the edge {a,b}\{a,b\} (or p=p=\infty if there is no edge). For instance, A(2)2A(2)\simeq\mathbb{Z}^{2} and A()𝔽2A(\infty)\simeq\mathbb{F}_{2}.

If aa and bb are different elements of SS, then the subgroup of A(Γ)A(\Gamma) spanned by aa and bb is isomorphic to the dihedral group A(p)A(p), where pp is the label of the edge {a,b}\{a,b\}. If p{2,}p\not\in\{2,\infty\}, the center of A(p)A(p) is the infinite cyclic group spanned by zab=wq(a,b)z_{ab}=w_{q}(a,b), where q=2pq=2p if pp is odd, and q=pq=p if pp is even (see [BS72] and [Del72] for the center of spherical Artin groups). If p=2p=2, the dihedral group A(2)A(2) is free abelian, we will denote zab=abz_{ab}=ab in this case.

1.2 CAT(0) cube complexes

A finite dimensional cube complex XX is naturally endowed with two natural distances, defined piecewise on cubes: the L1L^{1} distance d1d_{1} and the L2L^{2} distance d2d_{2} (each edge has length 11). Throughout the paper, unless we want to use both distances, we will mainly use the L1L^{1} distance d1d_{1} and will simply denote it dd.

A cube complex XX is called CAT(0) if the d2d_{2} distance is CAT(0), or equivalently if the d1d_{1} distance is median (see section 1.3). A discrete group GG is called cocompactly cubulated if it acts geometrically, i.e. properly and cocompactly by cubical isometries, on a CAT(0) cube complex.

Let us recall the fundamental local-to-global property for CAT(0) spaces.

Theorem 1.1 (Cartan-Hadamard).

A metric space is CAT(0) if and only if it is simply connected and locally CAT(0).

Let us recall Gromov’s combinatorial criterion to show that a cube complex is locally CAT(0).

Theorem 1.2 (Gromov, see [Gro87]).

A cube complex XX is locally CAT(0) if and only if, for any 33 cubes QQ,QQ^{\prime},Q′′Q^{\prime\prime} of XX, which pairwise intersect in codimension 11 and intersect globally in codimension 22, they are codimension 11 faces of some cube of XX.

In a CAT(0) cube complex XX, a hyperplane HH denotes the orthogonal (with respect to the CAT(0) metric d2d_{2}) of some edge [x,y][x,y] at its midpoint (i.e. the set of points of XX whose projection on [x,y][x,y] equals the midpoint), we denote it H=[x,y]H=[x,y]^{\perp} (the hyperplane HH can also be described as a union of midcubes, or as an equivalence class of edges, see [Sag95]). Each hyperplane divides XX into two connected components, the closures of which are called half-spaces and denoted by H+H^{+} and HH^{-}. An automorphism gg of XX is said to skewer the half-space H+H^{+} if gH+H+g\cdot H^{+}\subsetneq H^{+}. By skewering HH, we mean skewering H+H^{+} or HH^{-}.

We say that two hyperplanes H,HH,H^{\prime} cross if H+H^{+} and HH^{-} intersect H+H^{\prime+} and HH^{\prime-}.

If x,yx,y are vertices of a CAT(0) cube complex XX, then d1(x,y)d_{1}(x,y), also called the combinatorial distance between xx and yy, coincides with the number of hyperplanes separating xx and yy. An automorphism gg of XX is called combinatorially hyperbolic if gg preserves a combinatorial (d1d_{1}) geodesic, on which it acts by a nontrivial translation. An action of a group GG by cubical automorphisms on XX is called combinatorially semisimple if every element of GG is either combinatorially hyperbolic or fixes a vertex.

If XX is a cube complex, we can divide naturally each dd-cube into 2d2^{d} smaller cubes, getting a new cube complex (up to rescaling the metric by 22) called the cubical subdivision of XX.

Theorem 1.3 (Haglund, see [Hag07]).

Let GG be a group acting by cubical automorphisms on a CAT(0) cube complex XX. Then GG acts combinatorially semisimply on the cubical subdivision of XX.

In particular, if GG acts properly on XX and gGg\in G has infinite order, then gg acts as a combinatorial hyperbolic isometry of the cubical subdivision of XX.

If gg is a cubical isometry of a CAT(0) cube complex, its (combinatorial) translation length is

δg=minxX(0)d1(x,gx).\delta_{g}=\min_{x\in X^{(0)}}d_{1}(x,g\cdot x).

If AA is a subgroup of GG, its (combinatorial) minimal set is

Min(A)={xX|aA,d1(x,ax)=δa}.\operatorname{Min}(A)=\{x\in X\,|\,\forall a\in A,d_{1}(x,a\cdot x)=\delta_{a}\}.

If gGg\in G, we will simply denote Min(g)\operatorname{Min}(g) instead of Min(g)\operatorname{Min}(\langle g\rangle).

Remark.

Note that, according to Theorem 1.3, for any cubical automorphism gg of a CAT(0) cube complex XX, we have Min(g)\operatorname{Min}(g)\neq\emptyset. Furthermore, up to passing to the cubical subdivision of XX, we have Min(0)(g)=Min(g)X(0)\operatorname{Min}^{(0)}(g)=\operatorname{Min}(g)\cap X^{(0)}\neq\emptyset.

Remark.

Also note that Min(g)\operatorname{Min}(g) need not be convex for the d1d_{1} distance, nor need it be a cube subcomplex: consider for instance X=2X=\mathbb{R}^{2}, with the standard Cayley square complex structure of 2\mathbb{Z}^{2}, and let g:(x,y)(y+1,x+1)g:(x,y)\mapsto(y+1,x+1). Then δg=2\delta_{g}=2 and Min(g)={(x,y)2||xy|1}\operatorname{Min}(g)=\{(x,y)\in\mathbb{R}^{2}\,|\,|x-y|\leqslant 1\} is not a cube subcomplex.

We can also give a slight variation of this result, by considering an appropriate power of gg instead of considering the cubical subdivision.

Proposition 1.4.

Let XX be a CAT(0) cube complex of dimension at most DD, and let gg denote a cubical automorphism of XX. Then Min(g2D!)X(0)\operatorname{Min}(g^{2D!})\cap X^{(0)}\neq\emptyset.

Proof.

Let XX^{\prime} denote the cubical subidivision of XX. According to Theorem 1.3, either gg fixes a vertex of XX^{\prime} or gg is a combinatorially hyperbolic isometry of XX^{\prime}.

Assume that gg fixes a vertex of XX^{\prime}, i.e. gg stabilizes a cube QQ of XX. If QQ is a kk-cube, the isometry group of QQ is isomorphic to 𝔖k(/2)k\mathfrak{S}_{k}\ltimes\left(\mathbb{Z}/2\mathbb{Z}\right)^{k}. Each isometry of QQ has an order which divides 2k!2k!, and so it also divides 2D!2D!. Hence g2D!g^{2D!} fixes each vertex of QX(0)Q\cap X^{(0)}.

Assume that gg is a combinatorially hyperbolic isometry of XX^{\prime}. So there exists a vertex vv of XX^{\prime} such that {gnv,n}\{g^{n}\cdot v,n\in\mathbb{Z}\} lie on a combinatorial geodesic in XX^{\prime}. The vertex vv of XX^{\prime} corresponds to the midpoint of a cube QQ of XX. Choose vv such that the dimension kk of QQ is minimal. By minimality of kk, each hyperplane HH of XX (seen as a union of midcubes) containing vv also contains gvg\cdot v. So gg preserves the set of kk hyperplanes of XX containing vv. As a consequence, the element gD!g^{D!} fixes each hyperplane of XX containing vv, and g2D!g^{2D!} furthermore preserves the orientations of each of these kk hyperplanes. This implies that for each vertex xx of QX(0)Q\cap X^{(0)}, we have d(x,g2D!x)=d(v,g2D!v)d(x,g^{2D!}\cdot x)=d(v,g^{2D!}\cdot v). Hence xMin(g2D!)x\in\operatorname{Min}(g^{2D!}), so QX(0)Min(g2D!)Q\cap X^{(0)}\subset\operatorname{Min}(g^{2D!}). ∎

If XX is a proper CAT(0) cube complex, we will denote by X\partial_{\infty}X its visual (CAT(0)) boundary at infinity: it is endowed with the visual distance \sphericalangle. Each combinatorially hyperbolic isometry gg of XX is CAT(0) hyperbolic, and has a unique attracting fixed point g(+)Xg(+\infty)\in\partial_{\infty}X. See [BH99] for details on general CAT(0) spaces and isometries.

1.3 Median algebras

A median algebra is a set MM endowed with a symmetric map μ:M3M\mu:M^{3}\rightarrow M, called the median, satisfying the following

a,bM,μ(a,a,b)=a\displaystyle\forall a,b\in M,\mu(a,a,b)=a
a,b,c,d,eM,μ(a,b,μ(c,d,e))=μ(μ(a,b,c),μ(a,b,d),e).\displaystyle\forall a,b,c,d,e\in M,\mu(a,b,\mu(c,d,e))=\mu(\mu(a,b,c),\mu(a,b,d),e). (1)

In a metric space MM, the interval between aMa\in M and bMb\in M denotes I(a,b)={cM|d(a,c)+d(c,b)=d(a,b)}I(a,b)=\{c\in M\,|\,d(a,c)+d(c,b)=d(a,b)\}. A metric space MM is called metric median if

a,b,cM,I(a,b)I(b,c)I(a,c)={μ(a,b,c)},\forall a,b,c\in M,I(a,b)\cap I(b,c)\cap I(a,c)=\{\mu(a,b,c)\},

which implies that μ\mu is a median (see [BH83]).

Median algebras and CAT(0) cube complexes are highly related, as proved by Chepoi.

Theorem 1.5 ([Che00]).

A connected graph, endowed with its combinatorial distance, is metric median if and only if it is the 11-skeleton of a CAT(0) cube complex.

Recall that the rank of a subset II of a median algebra MM is the supremum of all nn\in\mathbb{N} such that the nn-cube {0,1}n\{0,1\}^{n} has a median-preserving embedding into II. Starting with a more general median algebra, one has the following.

Theorem 1.6 ([Nic04] and [CN05]).

Let MM be a median algebra with intervals of rank at most DD, and let GG be a group of automorphisms of MM. There exists a CAT(0) cube complex X(M)X(M), with vertex set X(M)(0)=MX(M)^{(0)}=M, of dimension at most DD, on which GG acts as a group of cubical automorphisms.

If XX is a CAT(0) cube complex and x,yXx,y\in X, let I(x,y)={zX|d1(x,z)+d1(z,y)=d1(x,y)}I(x,y)=\{z\in X\,|\,d_{1}(x,z)+d_{1}(z,y)=d_{1}(x,y)\} denote the d1d_{1} interval between xx and yy. A subset YXY\subset X is said to be convex if for every x,yYx,y\in Y, we have I(x,y)YI(x,y)\subset Y. If YXY\subset X, the d1d_{1} convex hull of YY is the smallest convex subset of XX containing YY, denoted Hull(Y)\operatorname{Hull}(Y).

The median μ:X3X\mu:X^{3}\rightarrow X is defined by

x,y,zX,I(x,y)I(y,z)I(z,x)={μ(x,y,z)}.\forall x,y,z\in X,I(x,y)\cap I(y,z)\cap I(z,x)=\{\mu(x,y,z)\}.

It is well-known that μ\mu is 11-Lipschitz for d1d_{1} with respect to each of the three variables. For the convenience of the reader, we give a short proof when restricted to the set of vertices.

Lemma 1.7.

The map μ:X(0)3X(0)\mu:{X^{(0)}}^{3}\rightarrow X^{(0)} is 11-Lipschitz with respect to the three variables, for the distance d1d_{1}.

Proof.

Let x,xX(0)x,x^{\prime}\in X^{(0)} be adjacent vertices of XX and y,zX(0)y,z\in X^{(0)}. Let m=μ(x,y,z)m=\mu(x,y,z) and m=μ(x,y,z)m^{\prime}=\mu(x^{\prime},y,z). Assume that HH is a hyperplane of XX separating mm and mm^{\prime}, say mH+m\in H^{+} and mHm^{\prime}\in H^{-}. Since m,mI(y,z)m,m^{\prime}\in I(y,z) we deduce that xH+x\in H^{+} and xHx^{\prime}\in H^{-}. So there exists at most 11 hyperplane separating mm and mm^{\prime}: d1(m,m)1d_{1}(m,m^{\prime})\leqslant 1. This implies that μ\mu is 11-Lipschitz. ∎

If II\subset\mathbb{R} is an interval, a map c:IXc:I\rightarrow X is called monotone if

stuI,μ(c(s),c(t),c(u))=c(t).\forall s\leqslant t\leqslant u\in I,\mu(c(s),c(t),c(u))=c(t).

If CXC\subset X is a non-empty convex subset, there exists a unique map πC:XC\pi_{C}:X\rightarrow C, called the gate projection onto CC (see for instance [BHS14]), such that

xX,cC,μ(x,πC(x),c)=πC(x).\forall x\in X,\forall c\in C,\mu(x,\pi_{C}(x),c)=\pi_{C}(x).

We now state a Lemma which will be used later.

Lemma 1.8.

Let XX denote a CAT(0) cube complex, and consider the median algebra (X(0),μ)(X^{(0)},\mu). Let MX(0)M\subset X^{(0)} denote a median subalgebra, and let CXC\subset X denote a convex subset. Then MCM\cap C is a convex subset of MM.

Proof.

Fix x,yMCx,y\in M\cap C, we want to prove that the interval IM(x,y)I_{M}(x,y) between xx and yy in the median algebra MM is contained in MCM\cap C. Let IX(x,y)I_{X}(x,y) denote the interval between xx and yy in XX, we have IM(x,y)=IX(x,y)MI_{M}(x,y)=I_{X}(x,y)\cap M. Since CC is convex in XX, we have IX(x,y)CI_{X}(x,y)\subset C. Hence IM(x,y)CMI_{M}(x,y)\subset C\cap M. So MCM\cap C is a convex subset of MM. ∎

If a group GG acts by cubical isometries on a CAT(0) cube complex XX, the action is said to be median minimal if X(0)X^{(0)} is the smallest GG-invariant non-empty median subalgebra of X(0)X^{(0)}.

2 Cubulation of centralizers

In this section, we prove the following result on cubulation of centralizers, which is more precise than Theorem I stated in the introduction.

Theorem 2.1.

Let GG be a group acting geometrically by isometries on a locally finite CAT(0) cube complex XX of dimension at most DD. Let AA be an abelian subgroup of GG such that every element of AA is the 2D!th2D!^{\text{th}} power of a combinatorially hyperbolic isometry in GG. Then the normalizer NG(A)N_{G}(A) of AA acts geometrically on the locally finite CAT(0) cube complex X(Min(A)(0))X(\operatorname{Min}(A)^{(0)}) of dimension at most DD associated to the median subalgebra Min(A)(0)\operatorname{Min}(A)^{(0)}. Furthermore, the centralizer ZG(A)Z_{G}(A) has finite index in NG(A)N_{G}(A).

Remark.

It is not always true that ZG(g)Z_{G}(g) acts cocompactly on Min(g2D!)(0)\operatorname{Min}(g^{2D!})^{(0)}: consider for instance X=2X=\mathbb{R}^{2}, with the standard Cayley square complex structure of 2\mathbb{Z}^{2}, and let g:(x,y)(y+1,x+1)g:(x,y)\mapsto(y+1,x+1) and h:(x,y)(x+1,y)h:(x,y)\mapsto(x+1,y). Consider the group GG spanned by gg and hh. We have ZG(g)=gZ_{G}(g)=\langle g\rangle\simeq\mathbb{Z}, but Min(g4)=2\operatorname{Min}(g^{4})=\mathbb{R}^{2}.

Lemma 2.2.

Let XX be a CAT(0) cube complex of dimension at most DD, and let gg be a combinatorially hyperbolic isometry of XX. Then for any xMin(g)(0)x\in\operatorname{Min}(g)^{(0)}, and for any hyperplane HH separating xx and gD!xg^{D!}\cdot x, gD!g^{D!} skewers HH.

Proof.

Fix xMin(g)(0)x\in\operatorname{Min}(g)^{(0)}, and let HH be a hyperplane such that xHx\in H^{-} and gD!xH+g^{D!}\cdot x\in H^{+}. Note that, since there is a combinatorial geodesic from xx to g2D!xg^{2D!}\cdot x through gD!xg^{D!}\cdot x crossing HH and gD!Hg^{D!}\cdot H, we know that gD!HHg^{D!}\cdot H\neq H.

Assume that for every 0i<jD0\leqslant i<j\leqslant D we have giHgjHg^{i}\cdot H\cap g^{j}\cdot H\neq\emptyset. Since gD!HHg^{D!}\cdot H\neq H, for every 0i<jD0\leqslant i<j\leqslant D, we have giHgjHg^{i}\cdot H\neq g^{j}\cdot H and giHgjHg^{i}\cdot H\cap g^{j}\cdot H\neq\emptyset so giHg^{i}\cdot H and gjHg^{j}\cdot H cross. Hence the D+1D+1 hyperplanes H,,gDHH,\ldots,g^{D}\cdot H pairwise cross, which is impossible in the cube complex XX with dimension at most DD.

As a consequence, there exist 0i<jD0\leqslant i<j\leqslant D such that giHgjH=g^{i}\cdot H\cap g^{j}\cdot H=\emptyset. Let k=jiDk=j-i\leqslant D, we have HgkH=H\cap g^{k}\cdot H=\emptyset. In particular, we have either H+gkH+H^{+}\subset g^{k}\cdot H^{+} or gkH+H+g^{k}\cdot H^{+}\subset H^{+}. Since kk divides D!D!, we have either H+gD!H+H^{+}\subset g^{D!}\cdot H^{+} or gD!H+H+g^{D!}\cdot H^{+}\subset H^{+}.

As gD!xgD!H+g^{D!}\cdot x\not\in g^{D!}\cdot H^{+} and gD!xH+g^{D!}\cdot x\in H^{+}, we conclude that gD!H+H+g^{D!}\cdot H^{+}\subset H^{+}. ∎

We now prove a very similar statement, but with the weaker assumption that xMin(gD!)(0)x\in\operatorname{Min}(g^{D!})^{(0)} instead of xMin(g)(0)x\in\operatorname{Min}(g)^{(0)}.

Lemma 2.3.

Let XX be a CAT(0) cube complex of dimension at most DD, and let gg be a combinatorially hyperbolic isometry of XX. Then for any xMin(gD!)(0)x\in\operatorname{Min}(g^{D!})^{(0)}, and for any hyperplane HH separating xx and gD!xg^{D!}\cdot x, gD!g^{D!} skewers HH.

Proof.

Fix xMin(gD!)(0)x\in\operatorname{Min}(g^{D!})^{(0)} and yMin(g)(0)y\in\operatorname{Min}(g)^{(0)}. By contradiction, assume that there exists a hyperplane HH such that xHx\in H^{-}, gD!xH+g^{D!}\cdot x\in H^{+} and HH is not skewered by gD!g^{D!}. According to Lemma 2.2, for every n,mn,m\in\mathbb{Z}, HH does not separate gnD!yg^{nD!}\cdot y and gmD!yg^{mD!}\cdot y. By symmetry, assume that n\forall n\in\mathbb{Z}, gnD!yH+g^{nD!}\cdot y\in H^{+}.

As a consequence, for every n0n\geqslant 0, gnD!Hg^{nD!}\cdot H separates xx and yy. Since only d1(x,y)d_{1}(x,y) hyperplanes separate xx and yy, we deduce that there exists n>0n>0 such that gnD!H=Hg^{nD!}\cdot H=H. This contradicts the fact that a combinatorial geodesic from xx to gnD!xg^{nD!}\cdot x goes via gD!xg^{D!}\cdot x as xMin(gD!)(0)x\in\operatorname{Min}(g^{D!})^{(0)} and crosses HH, whereas gnD!H=Hg^{nD!}\cdot H=H does not separate xx and gnD!xg^{nD!}\cdot x. ∎

Lemma 2.4.

Let XX be a CAT(0) cube complex of dimension at most DD, and let gg be a combinatorially hyperbolic isometry of XX of translation length δ\delta. Then the set

{hyperplanes of X skewered by gD!}/<gD!>\{\mbox{hyperplanes of $X$ skewered by $g^{D!}$}\}/<g^{D!}>

has cardinality D!δD!\delta.

Proof.

Let h=gD!h=g^{D!}, and fix xMin(h)(0)x\in\operatorname{Min}(h)^{(0)}.

Let HH be a hyperplane skewered by hh, so that hH+H+h\cdot H^{+}\subset H^{+}. Choose nn\in\mathbb{Z} such that d1(x,hnH˙)d_{1}(x,h^{n}\dot{H}) is minimal. Without loss of generality, assume that xhnH+x\in h^{n}\cdot H^{+}. Since hn+1Hh^{n+1}\cdot H is disjoint from hnHh^{n}\cdot H and hn+1H+hnH+h^{n+1}\cdot H^{+}\subset h^{n}\cdot H^{+}, we deduce that hn+1Hh^{n+1}\cdot H does not separate hnHh^{n}\cdot H and xx, and so hn+1Hh^{n+1}\cdot H separates xx and hxh\cdot x.

The number of hyperplanes separating xx and hxh\cdot x is equal to D!δD!\delta, so the cardinality of {\{hyperplanes of XX skewered by h}/<h>h\}/<h> is at most D!δD!\delta.

Fix a hyperplane HH separating xx and hxh\cdot x: according to Lemma 2.3, hh skewers HH. For instance, xHx\in H^{-}, hxH+h\cdot x\in H^{+} and hH+H+h\cdot H^{+}\subset H^{+}. As a consequence, if n>0n>0 then x,hxhnHx,h\cdot x\in h^{n}\cdot H^{-} so hnHh^{n}\cdot H does not separate xx and hxh\cdot x. Similarly, if n<0n<0 then x,hxhnH+x,h\cdot x\in h^{n}\cdot H^{+} so hnHh^{n}\cdot H does not separate xx and hxh\cdot x.

We conclude that the D!δD!\delta hyperplanes separating xx and hxh\cdot x are disjoint <h><h>-orbits, hence the cardinality of {\{hyperplanes of XX skewered by h}/<h>h\}/<h> is exactly D!δD!\delta. ∎

Proposition 2.5.

Let XX be a CAT(0) cube complex of dimension at most DD, and let gg be a combinatorially hyperbolic isometry of XX. Then Min(gD!)(0)\operatorname{Min}(g^{D!})^{(0)} is a median subalgebra of X(0)X^{(0)}, i.e. it is stable under the median of X(0)X^{(0)}.

Proof.

Let h=gD!h=g^{D!}, and let δ\delta denote the translation length of gg. Let μ\mu denote the median of XX.

Let x,y,zMin(h)(0)x,y,z\in\operatorname{Min}(h)^{(0)}, and let m=μ(x,y,z)X(0)m=\mu(x,y,z)\in X^{(0)}. Let HH be a hyperplane separating mm and hmh\cdot m, for instance mHm\in H^{-} and hmH+h\cdot m\in H^{+}. Since m=μ(x,y,z)Hm=\mu(x,y,z)\in H^{-} which is convex, at least two vertices among xx, yy and zz belong to HH^{-}: we can assume that x,yHx,y\in H^{-}. Similarly, hm=μ(hx,hy,hz)H+h\cdot m=\mu(h\cdot x,h\cdot y,h\cdot z)\in H^{+} which is convex, at least two vertices among hxh\cdot x, hyh\cdot y and hzh\cdot z belong to H+H^{+}: we can assume that hx,hzH+h\cdot x,h\cdot z\in H^{+}. As a consequence, HH separates xx and hxh\cdot x, so by Lemma 2.3 HH is skewered by hh.

According to Lemma 2.4, we conclude that at most D!δD!\delta hyperplanes separate mm and hmh\cdot m. Since the translation length of hh is D!δD!\delta, we conclude that d(m,hm)=D!δd(m,h\cdot m)=D!\delta, so mMin(h)(0)m\in\operatorname{Min}(h)^{(0)}. ∎

Remark.

There exists a combinatorially hyperbolic isometry gg of a locally finite CAT(0) cube complex such that for any n1n\geqslant 1, Min(gn)\operatorname{Min}(g^{n}) is not convex.

For instance, consider a infinite, rooted, binary tree, where each edge is replaced by the diagonal of a square: this defines a CAT(0) square complex TT. Then consider the isometry gg of TT defined recursively as shown in Figure 3. Notice that the fixed point set of g2g^{2} in TT is equal to the union of the two diagonals of the squares adjacent to the root, and it is not convex for the combinatorial distance d1d_{1}. Furthermore, for any nn\in\mathbb{N}, g2ng^{2^{n}} fixes some vertex vv (in fact, any vertex at the level nn) of TT and acts on the subcomplex TvT_{v} (corresponding to the subtree defined by vv) as gg. We deduce that, for any n1n\geqslant 1, the fixed point set of g2ng^{2^{n}} is not convex (we could argue similarly that, for every even n1n\geqslant 1, the fixed point set of g2ng^{2n} is not convex).

If we want to find an example where gg is a combinatorially hyperbolic isometry, it is sufficient to consider the direct product T×T\times\mathbb{R}, where gg acts on \mathbb{R} by a unit translation.

gidg
Figure 3: A pathological isometry of a CAT(0) square complex.

We will now show that the minimal set of an abelian group is not empty.

Proposition 2.6.

Let XX be a CAT(0) cube complex of dimension at most DD, and let AA be a finitely generated abelian group of cubical automorphisms of XX. Then Min(A2D!)(0)=aAMin(a2D!)(0)\operatorname{Min}(A^{2D!})^{(0)}=\bigcap_{a\in A}\operatorname{Min}(a^{2D!})^{(0)} is not empty.

Proof.

We will prove the result by induction on the numbers of generators of AA. When AA is cyclic it follows from Proposition 1.4. Assume that A=B,gA=\langle B,g\rangle, where Min(B2D!)(0)\operatorname{Min}(B^{2D!})^{(0)}\neq\emptyset.

Since gg commutes with B2D!B^{2D!}, gg preserves Min(B2D!)(0)\operatorname{Min}(B^{2D!})^{(0)}. According to Proposition 2.5, Min(B2D!)(0)\operatorname{Min}(B^{2D!})^{(0)} is a median subalgebra of X(0)X^{(0)}. The element gg acts as a cubical automorphism on the CAT(0) cube complex Y=X(Min(B2D!)(0))Y=X(\operatorname{Min}(B^{2D!})^{(0)}). Since Min(B2D!)(0)\operatorname{Min}(B^{2D!})^{(0)} is a median subalgebra of X(0)X^{(0)} which has rank at most DD, we deduce that YY has dimension at most DD.

According to Proposition 1.4, there exists a vertex xY(0)=Min(B2D!)(0)x\in Y^{(0)}=\operatorname{Min}(B^{2D!})^{(0)} which lies in MinY(g2D!)\operatorname{Min}_{Y}(g^{2D!}) (the minimal set of g2D!g^{2D!} considered as an isometry of YY). This means that {g2nD!x,n}\{g^{2nD!}\cdot x,n\in\mathbb{Z}\} lie on a combinatorial geodesic in YY. Since Min(B2D!)(0)\operatorname{Min}(B^{2D!})^{(0)} is a median subalgebra of X(0)X^{(0)}, we deduce that {g2nD!x,n}\{g^{2nD!}\cdot x,n\in\mathbb{Z}\} lie on a combinatorial geodesic in XX. This precisely means that xMin(g2D!)x\in\operatorname{Min}(g^{2D!}). This vertex xx of XX belongs to Min(g2D!)Min(B2D!)=Min(A2D!)\operatorname{Min}(g^{2D!})\cap\operatorname{Min}(B^{2D!})=\operatorname{Min}(A^{2D!}), which proves that Min(A2D!)(0)\operatorname{Min}(A^{2D!})^{(0)}\neq\emptyset. ∎

Proposition 2.7.

Let GG be a group acting geometrically on a locally finite CAT(0) cube complex XX of dimension at most DD, and let AA be an abelian subgroup of GG consisting of 2D!th2D!^{\text{th}} powers of combinatorially hyperbolic isometries. Then the centralizer ZG(A)Z_{G}(A) of AA in GG has finite index in the normalizer NG(A)N_{G}(A) of AA in GG, and NG(A)N_{G}(A) acts geometrically on Min(A)(0)\operatorname{Min}(A)^{(0)}.

Proof.

According to Proposition 2.6, the minimal set Min(A)(0)\operatorname{Min}(A)^{(0)} is not empty.

The action of NG(A)N_{G}(A) on XX is proper and stabilizes Min(A)(0)\operatorname{Min}(A)^{(0)}, so it induces a proper action on Min(A)(0)\operatorname{Min}(A)^{(0)}.

Assume that the action of ZG(A)Z_{G}(A) on Min(A)(0)\operatorname{Min}(A)^{(0)} is not cocompact: since GG acts properly and cocompactly on XX, there exist C0C\geqslant 0, xMin(A)(0)x\in\operatorname{Min}(A)^{(0)} and (hn)nG(h_{n})_{n\in\mathbb{N}}\in G^{\mathbb{N}} such that n,d(hnx,Min(A)(0))C\forall n\in\mathbb{N},d(h_{n}\cdot x,\operatorname{Min}(A)^{(0)})\leqslant C and the cosets (hnZG(A))n(G/ZG(A))(h_{n}Z_{G}(A))_{n\in\mathbb{N}}\in\left(G/Z_{G}(A)\right)^{\mathbb{N}} are pairwise distinct.

According to the flat torus theorem (see [BH99, Theorem 7.1]), the abelian group AA acts properly by semisimple isometries on the CAT(0) space XX, so AA is finitely generated. Fix a1,,ara_{1},\dots,a_{r} some generators of AA. Fix some 1ir1\leqslant i\leqslant r.

Let δi\delta_{i} denote the combinatorial translation length of aia_{i}. For all nn\in\mathbb{N}, since d(hnx,Min(ai)(0))Cd(h_{n}\cdot x,\operatorname{Min}(a_{i})^{(0)})\leqslant C we have d(hnx,aihnx)δi+2Cd(h_{n}\cdot x,a_{i}h_{n}\cdot x)\leqslant\delta_{i}+2C. So d(x,hn1aihnx)δi+2Cd(x,h_{n}^{-1}a_{i}h_{n}\cdot x)\leqslant\delta_{i}+2C. Since XX is locally finite, up to passing to a subsequence, we may assume that n,m,hn1aihnx=hm1aihmx\forall n,m\in\mathbb{N},h_{n}^{-1}a_{i}h_{n}\cdot x=h_{m}^{-1}a_{i}h_{m}\cdot x. Since the action of GG on XX is proper, the stabilizer of xx is finite, so up to passing to a new subsequence, we may assume that n,m,hn1aihn=hm1aihm\forall n,m\in\mathbb{N},h_{n}^{-1}a_{i}h_{n}=h_{m}^{-1}a_{i}h_{m}. So n,m,hnhm1\forall n,m\in\mathbb{N},h_{n}h_{m}^{-1} centralizes aia_{i}.

If we apply this for every 1ir1\leqslant i\leqslant r, we obtain up to passing to a new subsequence that n,m,hnhm1\forall n,m\in\mathbb{N},h_{n}h_{m}^{-1} centralizes a1,,ara_{1},\dots,a_{r}. Since a1,,ara_{1},\dots,a_{r} span AA, we deduce that n,m,hnhm1ZG(A)\forall n,m\in\mathbb{N},h_{n}h_{m}^{-1}\in Z_{G}(A). This contradicts the assumption that the cosets (hnZG(A))n(G/ZG(A))(h_{n}Z_{G}(A))_{n\in\mathbb{N}}\in\left(G/Z_{G}(A)\right)^{\mathbb{N}} are pairwise distinct.

As a consequence, the induced action of ZG(A)Z_{G}(A) on Min(A)(0)\operatorname{Min}(A)^{(0)} is proper and cocompact. Since the action of NG(A)N_{G}(A) on Min(A)(0)\operatorname{Min}(A)^{(0)} is also proper, we deduce that ZG(A)Z_{G}(A) has finite index in NG(A)N_{G}(A). ∎

We obtain now the proof of Theorem 2.1.

Proof.

Let a1,,ara_{1},\dots,a_{r} be some generators of AA. For each 1ir1\leqslant i\leqslant r, by Proposition 2.5 Min(ai)(0)\operatorname{Min}(a_{i})^{(0)} is a median subalgebra of X(0)X^{(0)}. As a consequence, Min(A)(0)=i=1rMin(ai)(0)\operatorname{Min}(A)^{(0)}=\bigcap_{i=1}^{r}\operatorname{Min}(a_{i})^{(0)} is a also a median subalgebra of X(0)X^{(0)}. According to Proposition 2.6, it is not empty. By Proposition 2.7, NG(A)N_{G}(A) acts properly cocompactly on Min(A)(0)\operatorname{Min}(A)^{(0)}. Theorem 1.6 concludes the proof.

Note that the CAT(0) cube complex X(Min(A)(0))X(\operatorname{Min}(A)^{(0)}) has dimension at most DD, since Min(A)(0)\operatorname{Min}(A)^{(0)} is a median subalgebra of X(0)X^{(0)} which has rank at most DD. ∎

Remark.

Note that the distances induced on Min(A)(0)\operatorname{Min}(A)^{(0)} by XX and by X(Min(A)(0))X(\operatorname{Min}(A)^{(0)}) may be different.

For instance, consider the action of A=gA=\langle g\rangle\simeq\mathbb{Z} on X=2X=\mathbb{R}^{2} by g(x,y)=(y+1,x+1)g\cdot(x,y)=(y+1,x+1). Then Min(g)(0)={(n,n)|n}\operatorname{Min}(g)^{(0)}=\{(n,n)\,|\,n\in\mathbb{Z}\}. Inside XX, the combinatorial distance between (n,n)(n,n) and (n+1,n+1)(n+1,n+1) is equal to 22. But as a median algebra, Min(g)(0)\operatorname{Min}(g)^{(0)} is isomorphic to \mathbb{Z}, and hence X(Min(g)(0))X(\operatorname{Min}(g)^{(0)}) is isomorphic to the standard cubical structure on \mathbb{R}. So in X(Min(g)(0))X(\operatorname{Min}(g)^{(0)}), the combinatorial distance between (n,n)(n,n) and (n+1,n+1)(n+1,n+1) is equal to 11. This is because the two hyperplanes of 2\mathbb{R}^{2} separating (n,n)(n,n) and (n+1,n+1)(n+1,n+1) define the same partition of Min(g)(0)\operatorname{Min}(g)^{(0)}.

This example does not satisfy the assumption of Theorem 2.1 that elements of AA are fourth powers of isometries, but nevertheless it illustrates the difference between the two cubical structures on Min(A)(0)\operatorname{Min}(A)^{(0)}.

3 Convex-cocompact subgroups

In this article, the rank of a finitely generated abelian or virtually abelian group is the minimal number of generators of a finite index free abelian subgroup.

Definition 3.1.

A subgroup AA of a group GG acting geometrically on a CAT(0) cube complex XX is said to be convex-cocompact in XX if there exists a convex subcomplex YXY\subset X which is AA-invariant and AA-cocompact.

We now give an equivalent characterization of convex-cocompact subgroups, starting with a small Lemma.

Lemma 3.2.

Let XX be a finite-dimensional cube complex and let YXY\subset X be a convex subcomplex. For any nn\in\mathbb{N}, there exists a convex subcomplex YnY_{n} of XX containing the combinatorial neighbourhood of radius nn of YY, and such that YnY_{n} is contained in a bounded neighbourhood of YY. Furthermore, if gg is an automorphism of XX preserving YY, then gg preserves YnY_{n} for every nn\in\mathbb{N}.

Proof.

It is sufficient to prove the statement for n=1n=1. Let ZZ denote the full subcomplex of XX whose vertices are all vertices xX(0)x\in X^{(0)} such that any two hyperplanes separating xx and YY cross. Then ZZ contains the combinatorial 11-neighbourhood of YY, we will prove that ZZ is convex.

Fix z,zZ(0)z,z^{\prime}\in Z^{(0)}, and consider a vertex xx on a combinatorial geodesic from zz to zz^{\prime}, we will prove that xZx\in Z. Assume that there exist two distinct hyperplanes H,HH,H^{\prime} separating xx and YY. We will prove that HH and HH^{\prime} cross. Since xx lies on a combinatorial geodesic from zz to zz^{\prime}, we deduce that HH separates zz and YY, or separates zz^{\prime} and YY. We deduce the same for HH^{\prime}.

If both HH and HH^{\prime} separate zz and YY, then HH and HH^{\prime} cross since zZz\in Z.

So we can assume that HH separates {z,x}\{z,x\} and {z}Y\{z^{\prime}\}\cup Y, and that HH^{\prime} separates {z,x}\{z^{\prime},x\} and {z}Y\{z\}\cup Y. Hence the four intersections of the half-spaces of HH and HH^{\prime} are non-empty, so HH and HH^{\prime} cross.

As a consequence, ZZ is convex.

Finally, for any vertex zZ(0)z\in Z^{(0)}, let yY(0)y\in Y^{(0)} denote the gate projection of zz onto YY. Then by definition of ZZ, all hyperplanes separating zz and yy pairwise cross. As a consequence, zz and yy belong to a common cube of XX. We deduce that yy and zz are at distance at most the dimension of XX. In particular, ZZ is contained in a bounded neighbourhood of YY. ∎

Proposition 3.3.

Let AA be a subgroup of a group GG acting geometrically on a CAT(0) cube complex XX. Then AA is convex-cocompact if and only if for every xX(0)x\in X^{(0)} (equivalently, for some xX(0)x\in X^{(0)}), AA acts cocompactly on Hull(Ax)\operatorname{Hull}(A\cdot x).

Proof.

If AA is convex-cocompact, for any xYx\in Y, we have Hull(Ax)Y\operatorname{Hull}(A\cdot x)\subset Y and so AA acts cocompactly on Hull(Ax)\operatorname{Hull}(A\cdot x).

Conversely, assume that there exists a vertex xX(0)x\in X^{(0)} such that AA acts cocompactly on Y=Hull(Ax)Y=\operatorname{Hull}(A\cdot x). For any vertex yX(0)y\in X^{(0)}, according to Lemma 3.2, there exists a convex subcomplex ZZ of XX containing ZZ and yy, contained in a neighbourhood of YY, which is furthermore AA-invariant. Since XX is locally finite, we deduce that AA acts cocompactly on ZZ. Since Hull(Ay)Z\operatorname{Hull}(A\cdot y)\subset Z, we conclude that AA acts cocompactly on Hull(Ay)\operatorname{Hull}(A\cdot y). ∎

Remark.

Note that being convex-cocompact depends on the CAT(0) cube complex XX: see for instance Subsection 5.2.

We will now state a few technical results about convex-cocompact subgroups which will be used in the sequel.

Lemma 3.4.

Let GG be a group acting geometrically on a CAT(0) cube complex XX, and let A,BA,B be subgroups of GG which are convex-cocompact in XX. Then ABA\cap B is convex-cocompact in XX.

Proof.

Fix a vertex xX(0)x\in X^{(0)}, and consider a sequence (xn)n(x_{n})_{n\in\mathbb{N}} in Hull(ABx)\operatorname{Hull}(A\cap B\cdot x). According to Proposition 3.3, AA and BB act cocompactly on Hull(Ax)\operatorname{Hull}(A\cdot x) and Hull(Bx)\operatorname{Hull}(B\cdot x) respectively, so there exist C>0C>0 and sequences (an)n(a_{n})_{n\in\mathbb{N}} in AA and (bn)n(b_{n})_{n\in\mathbb{N}} in BB such that n,d(anx,xn)C\forall n\in\mathbb{N},d(a_{n}\cdot x,x_{n})\leqslant C and d(bnx,xn)Cd(b_{n}\cdot x,x_{n})\leqslant C. As a consequence, n,d(bn1anx,x)2C\forall n\in\mathbb{N},d(b_{n}^{-1}a_{n}\cdot x,x)\leqslant 2C. Since GG acts properly on the CAT(0) cube complex, we deduce that, up to passing to subsequences, we have n,m,bm1am=bn1an\forall n,m\in\mathbb{N},b_{m}^{-1}a_{m}=b_{n}^{-1}a_{n}, so anam1=bnbm1ABa_{n}a_{m}^{-1}=b_{n}b_{m}^{-1}\in A\cap B. So for all nn\in\mathbb{N}, we have d(ana01x,xn)d(anx,xn)+d(ana01x,anx)C+d(a01x,x)d(a_{n}a_{0}^{-1}\cdot x,x_{n})\leqslant d(a_{n}\cdot x,x_{n})+d(a_{n}a_{0}^{-1}\cdot x,a_{n}\cdot x)\leqslant C+d(a_{0}^{-1}\cdot x,x) is bounded. Since n,ana01AB\forall n\in\mathbb{N},a_{n}a_{0}^{-1}\in A\cap B, this proves that ABA\cap B acts cocompactly on Hull(ABx)\operatorname{Hull}(A\cap B\cdot x). Hence ABA\cap B is convex-cocompact in XX. ∎

Definition 3.5.

A virtually abelian subgroup AA of a group GG is called highest if for any virtually abelian subgroup BB of GG such that ABA\cap B has finite index in AA, we have that ABA\cap B has finite index in BB.

We now recall the following recent result from D. Wise and D. Woodhouse.

Theorem 3.6 (Cubical flat torus theorem [WW17]).

Let GG be a group acting geometrically on a CAT(0) cube complex XX. Let AA be a highest virtually abelian subgroup of GG. Then AA is convex-cocompact in XX.

We can now prove the following, which is the main technical part in the proof of Theorem J stated in the introduction.

Lemma 3.7.

Let GG be a group acting geometrically, combinatorially semisimply by isometries on a CAT(0) cube complex XX, median minimally. Let WW be a central subgroup of GG which is convex-cocompact in XX. Then XX splits as a product of two convex cube subcomplexes XY×ZX\simeq Y\times Z, where GG preserves this splitting, WW acts with finite index kernel WW^{\prime} on YY and G/WG/W^{\prime} acts geometrically on YY, with dimYdimXrkW\operatorname{dim}Y\leqslant\operatorname{dim}X-\operatorname{rk}W. Furthermore, if AGA\subset G is convex-cocompact in XX, then AW/WAW^{\prime}/W^{\prime} is convex-cocompact in YY.

Proof.

Let DD denote the dimension of XX, and let W=W2D!2=w2D!2|wWW^{\prime}=W^{2D!^{2}}=\langle w^{2D!^{2}}\,|\,w\in W\rangle. Since WW is abelian and finitely generated, WW^{\prime} has finite index in WW.

Let Skew(W)Skew(W^{\prime}) denote the set of hyperplanes of XX that are skewered by at least one element of WW^{\prime}. Let Stab(W)Stab(W^{\prime}) denote the set of hyperplanes of XX that are stabilized by all elements of WW^{\prime}. We will prove that the set of hyperplanes of XX is the disjoint GG-invariant union Skew(W)Stab(W)Skew(W^{\prime})\sqcup Stab(W^{\prime}). By definition, the sets Skew(W)Skew(W^{\prime}) and Stab(W)Stab(W^{\prime}) are disjoint. Since WW is central in GG, both sets Skew(W)Skew(W^{\prime}) and Stab(W)Stab(W^{\prime}) are GG-invariant.

Let us first remark that each hyperplane in Skew(W)Skew(W^{\prime}) intersects each hyperplane in Stab(W)Stab(W^{\prime}). Fix HSkew(W)H\in Skew(W^{\prime}), it is skewered by some wWw\in W^{\prime}. Fix HStab(W)H^{\prime}\in Stab(W^{\prime}), it is stabilized by ww. Since HH is skewered by ww, any w\langle w\rangle-orbit in XX intersects both half-spaces of HH. In particular, if xHx\in H^{\prime}, we know that wx\langle w\rangle\cdot x lies in HH^{\prime} and intersects both half-spaces of HH. Hence HH intersects HH^{\prime}.

So we have to prove that every hyperplane of XX is in the union Skew(W)Stab(W)Skew(W^{\prime})\sqcup Stab(W^{\prime}). Fix a hyperplane HH which is not in Skew(W)Skew(W^{\prime}), we will prove that HStab(W)H\in Stab(W^{\prime}).

Since WW is abelian, according to Proposition 2.6, we have Min(W2D!)(0)\operatorname{Min}(W^{2D!})^{(0)}\neq\emptyset. According to Proposition 2.5, Min(W2D!)(0)\operatorname{Min}(W^{2D!})^{(0)} is a median subalgebra of X(0)X^{(0)}. As W2D!W^{2D!} is normal in GG, we deduce that GG stabilizes Min(W2D!)(0)\operatorname{Min}(W^{2D!})^{(0)}. As the action of GG on X(0)X^{(0)} is median minimal, we deduce that X(0)=Min(W2D!)(0)X^{(0)}=\operatorname{Min}(W^{2D!})^{(0)}.

Note that any decreasing sequence of W2D!W^{2D!}-cocompact subcomplexes of XX terminates, so there exists a vertex xX(0)x\in X^{(0)} for which the action of W2D!W^{2D!} on Z=Hull(W2D!x)Z=\operatorname{Hull}(W^{2D!}\cdot x) is such that ZZ is the smallest non-empty convex W2D!W^{2D!}-invariant subcomplex of ZZ.

Any hyperplane intersecting ZZ separates two points in W2D!xW^{2D!}\cdot x, so according to Lemma 2.2 such a hyperplane is skewered by an element of W2D!W^{2D!}. Hence HH does not intersect ZZ.

Since the action of GG on XX is median minimal, the orbit GxG\cdot x is not contained in a single half-space of HH, so there exists gGg\in G such that HH separates xx and gxg\cdot x. Say ZHZ\subset H^{-} and gxH+g\cdot x\in H^{+}.

We will show that for every wW2D!w\in W^{2D!}, we have wD!H=Hw^{D!}\cdot H=H. Fix wW2D!{1}w\in W^{2D!}\char 92\relax\{1\}. We will first show that HH separates ZZ and gwxg\langle w\rangle\cdot x. So we want to show that gwxH+g\langle w\rangle\cdot x\subset H^{+}.

By contradiction, assume that there exists n00n_{0}\neq 0 such that gwn0xHgw^{n_{0}}\cdot x\in H^{-}. Without loss of generality (up to replacing ww with w1w^{-1}), we can assume that n0>0n_{0}>0.

As gxX(0)=Min(W2D!)(0)g\cdot x\in X^{(0)}=\operatorname{Min}(W^{2D!})^{(0)}, we deduce that (gwkx)k(gw^{k}\cdot x)_{k\in\mathbb{Z}} lies on a combinatorial geodesic of XX. As a consequence, for every n0n\leqslant 0, we have gwnxH+gw^{n}\cdot x\in H^{+}. So for every n0n\leqslant 0, we have gxwnH+g\cdot x\in w^{-n}\cdot H^{+} and xZ=wnZwnHx\in Z=w^{-n}\cdot Z\subset w^{-n}\cdot H^{-}.

Since finitely many hyperplanes separate xx and gxg\cdot x, we conclude that there exists n<0n<0 such that wnH=Hw^{n}\cdot H=H. On one hand, as |n|n0n0|n|n_{0}\geqslant n_{0} and (gwkx)k(gw^{k}\cdot x)_{k\in\mathbb{Z}} lies on a combinatorial geodesic of XX, we have gw|n|n0xHgw^{|n|n_{0}}\cdot x\in H^{-}. On the other hand, as w|n|n0H+=H+w^{|n|n_{0}}\cdot H^{+}=H^{+} and gxH+g\cdot x\in H^{+} we have gw|n|n0xH+gw^{|n|n_{0}}\cdot x\in H^{+}. This is a contradiction.

As a consequence, every hyperplane separating ZZ and gxg\cdot x separates ZZ and gwxg\langle w\rangle\cdot x. Let {\cal H} denote the finite set of hyperplanes separating ZZ and gxg\cdot x. We have seen that ww preserves the set {\cal H}, and acts as a bijection σ\sigma on {\cal H}. Since ww preserves ZZ, we know that hyperplanes of {\cal H} in the same σ\sigma-orbit are at the same distance of ZZ, and thus cannot be nested. Hence each kk-cycle of σ\sigma in {\cal H} corresponds to kk pairwise crossing hyperplanes, so kDk\leqslant D. As a consequence, σD!=1\sigma^{D!}=1. Since HH\in{\cal H}, we conclude that wD!H=Hw^{D!}\cdot H=H. Furthermore, since every element of GG acts combinatorially semisimply, we deduce that wD!w^{D!} stabilizes each half-space of HH.

We have proved that for every wW2D!w\in W^{2D!}, we have wD!H=Hw^{D!}\cdot H=H, so HStab(W)H\in Stab(W^{\prime}).

The GG-equivariant disjoint decomposition Skew(W)Stab(W)Skew(W^{\prime})\sqcup Stab(W^{\prime}) defines a GG-equivariant isomorphism XY×ZX\simeq Y\times Z, where YY is the cube complex dual to the set of hyperplanes Stab(W)Stab(W^{\prime}). Note that ZZ is also isomorphic to the cube complex dual to the set of hyperplanes Skew(W)Skew(W^{\prime}).

Since WW acts properly on ZZ, we have dimZrkW\operatorname{dim}Z\geqslant\operatorname{rk}W, so dimYdimXrkW\operatorname{dim}Y\leqslant\operatorname{dim}X-\operatorname{rk}W. Furthermore, by definition of Stab(W)Stab(W^{\prime}), we know that WW^{\prime} acts trivially on YY, so that the GG-action on YY factors through G/WG/W^{\prime}.

As GG acts cocompactly on XX, we deduce that G/WG/W^{\prime} acts cocompactly on YY. Since WW^{\prime} acts properly on ZZ, we deduce that G/WG/W^{\prime} acts properly on YY. In conclusion, G/WG/W^{\prime} acts geometrically on YY.

Furthermore, if AGA\subset G is convex-cocompact in XX, then AA acts cocompactly on a convex subcomplex MAM_{A} of XX. Since XY×ZX\simeq Y\times Z, the convex subcomplex MAM_{A} splits as MAMA,Y×MA,ZM_{A}\simeq M_{A,Y}\times M_{A,Z}, such that AW/WAW^{\prime}/W^{\prime} acts geometrically on the convex subcomplex MA,YM_{A,Y} of YY. ∎

Lemma 3.8.

Let GG be a group acting geometrically on a CAT(0) cube complex XX. Assume that HGH\subset G acts geometrically on a median subalgebra MX(0)M\subset X^{(0)}, with associated CAT(0) cube complex X(M)X(M). Assume that AA is a convex-cocompact subgroup of GG. Then AHA\cap H is convex-cocompact in X(M)X(M).

Proof.

Consider a convex subcomplex YY of XX such that AA acts properly and cocompactly on YY. Up to considering some convex neighbourhood of YY, we may assume that YY and MM intersect. Let x0YMx_{0}\in Y\cap M.

The group AHA\cap H acts properly on YMY\cap M. We will prove that the action is cocompact. By contradiction, assume that there exists a sequence (yn)n(y_{n})_{n\in\mathbb{N}} in YMY\cap M such that d1(yn,AHx0)n++d_{1}(y_{n},A\cap H\cdot x_{0})\underset{n\rightarrow{+\infty}}{\longrightarrow}{+\infty}. Since the action of GG on XX is cocompact, there exists D0D\geqslant 0 such that for each nn\in\mathbb{N}, there exists gnGg_{n}\in G such that d1(yn,gnx0)Dd_{1}(y_{n},g_{n}\cdot x_{0})\leqslant D, so that d1(gnx0,AHx0)n++d_{1}(g_{n}\cdot x_{0},A\cap H\cdot x_{0})\underset{n\rightarrow{+\infty}}{\longrightarrow}{+\infty}.

Since the action of AA on YY is cocompact and d1(gnx0,Y)Dd_{1}(g_{n}\cdot x_{0},Y)\leqslant D for all nn\in\mathbb{N}, there exists a finite subset KK of GG such that, for each nn\in\mathbb{N}, we have gnAKg_{n}\in AK. Similarly, since the action of HH on MM is cocompact and d1(gnx0,M)Dd_{1}(g_{n}\cdot x_{0},M)\leqslant D for all nn\in\mathbb{N}, there exists a finite subset KK^{\prime} of GG such that, for each nn\in\mathbb{N}, we have gnHKg_{n}\in HK^{\prime}. Up to passing to a subsequence, we may assume that there exists kKk\in K and kKk^{\prime}\in K^{\prime} such that for each nn\in\mathbb{N}, we have gnAkHkg_{n}\in Ak\cap Hk^{\prime}. In particular, for each nn\in\mathbb{N} we have gng01AHg_{n}g_{0}^{-1}\in A\cap H. So d1(gnx0,AHx0)=d1((gng01)g0x0,AHx0)=d1(g0x0,AHx0)d_{1}(g_{n}\cdot x_{0},A\cap H\cdot x_{0})=d_{1}((g_{n}g_{0}^{-1})g_{0}\cdot x_{0},A\cap H\cdot x_{0})=d_{1}(g_{0}\cdot x_{0},A\cap H\cdot x_{0}) is bounded, which is a contradiction.

So we have proved that the group AHA\cap H acts properly and cocompactly on YMY\cap M. Since Y(0)Y^{(0)} is convex in the median algebra X(0)X^{(0)}, according to Lemma 1.8, we deduce that MY(0)M\cap Y^{(0)} is convex in the median subalgebra MM. This implies that AHA\cap H is a convex-cocompact subgroup in X(M)X(M). ∎

4 Non-cubicality criterion

We will now summarize the main two stability results for virtual cubulation that we will use. They are slightly more precise than Theorem I (only the part about centralizers) and Theorem J.

Proposition 4.1.

Let GG be a group acting geometrically on a CAT(0) cube complex XX, and let CC be an abelian subgroup of GG. There exists a finite index subgroup C0C_{0} of CC such that the centralizer ZG(C0)Z_{G}(C_{0}) acts geometrically on a CAT(0) cube complex YY with dimYdimX\operatorname{dim}Y\leqslant\operatorname{dim}X.

Furthermore, if AGA\subset G is convex-cocompact in XX, then AZG(C0)A\cap Z_{G}(C_{0}) is convex-cocompact in YY.

Proof.

This is a consequence of Theorem 2.1 and Lemma 3.8. ∎

Proposition 4.2.

Let GG be a group acting geometrically on a CAT(0) cube complex XX, and let WW be a central subgroup of GG which is convex-cocompact in XX. Then WW has a finite index subgroup WW^{\prime} such that G/WG/W^{\prime} acts geometrically on a CAT(0) cube complex YY with dimYdimXrkW\operatorname{dim}Y\leqslant\operatorname{dim}X-\operatorname{rk}W.

Furthermore, if AGA\subset G is convex-cocompact in XX, then AW/WAW^{\prime}/W^{\prime} is convex-cocompact in YY.

Proof.

This is contained in Lemma 3.7. The combinatorially semisimple assumption follows from Theorem 1.3, up to passing to the cubical subdivision of XX. Furthermore, one can always assume that the action is median minimal, up to passing to a smaller median subalgebra. Note that this does not increase the dimension of XX: indeed if MM^{\prime} is a median subalgebra of MM, then cubes in MM^{\prime} are cubes in MM. ∎

We now give a slightly more general version of a result from Crisp and Paoluzzi (see [CP05]), which studies proper semisimple actions of B3B_{3} and B4B_{4} on CAT(0) spaces. Note that there is no cocompactness assumption in this result, nor a CAT(0) cube complex.

Proposition 4.3 (Crisp-Paoluzzi).

Let p3p\in\mathbb{N}_{\geqslant 3} and consider the dihedral Artin-Tits group A=A(p)=a,bwp(a,b)=wp(b,a)A=A(p)=\langle a,b\mid w_{p}(a,b)=w_{p}(b,a)\rangle. Assume AA acts properly, by semisimple isometries on a CAT(0) space XX. Then aa, zabz_{ab} and bb act by hyperbolic isometries, whose attracting endpoints in the visual boundary X\partial_{\infty}X are denoted a(+)a({+\infty}), zab(+)z_{ab}({+\infty}) and b(+)b({+\infty}). Furthermore, if we denote by \sphericalangle the visual distance on X\partial X, we have:

  • \bullet

    If pp is odd, then we have (a(+),zab(+))<π2\sphericalangle(a({+\infty}),z_{ab}({+\infty}))<\frac{\pi}{2} and (b(+),zab(+))<π2\sphericalangle(b({+\infty}),z_{ab}({+\infty}))<\frac{\pi}{2}.

  • \bullet

    If pp is even, then we have (a(+),zab(+))<π2\sphericalangle(a({+\infty}),z_{ab}({+\infty}))<\frac{\pi}{2} or (b(+),zab(+))<π2\sphericalangle(b({+\infty}),z_{ab}({+\infty}))<\frac{\pi}{2}.

Proof.

We adapt here the proof of [CP05, Theorem 4]. Without loss of generality, we may assume that the action of AA on XX is minimal. By properness, every infinite order element of AA acts by a hyperbolic isometry, in particular aa, bb and zabz_{ab}. Then by [BH99, Theorem II.6.8], XX is isometric to the product ×Y\mathbb{R}\times Y, where YY is a CAT(0) space, and Z(A)=zabZ(A)=\langle z_{ab}\rangle acts by translation on \mathbb{R} and trivially on YY. Let δ{0}\delta\in\mathbb{R}\char 92\relax\{0\} such that zabz_{ab} acts on the \mathbb{R} factor by a translation of δ\delta. Up to the choice of the orientation of \mathbb{R}, we may assume that δ>0\delta>0.

Since aa and bb commute with zabz_{ab}, they preserve the decomposition X×YX\simeq\mathbb{R}\times Y and preserve the orientation of the \mathbb{R} factor. In particular, let α,β\alpha,\beta\in\mathbb{R} such that aa and bb act on the \mathbb{R} factor by translations of α\alpha and β\beta respectively.

  • \bullet

    If pp is odd, then aa and bb are conjugated by wp(a,b)w_{p}(a,b) in AA, we deduce that α=β\alpha=\beta. But zab=w2p(a,b)z_{ab}=w_{2p}(a,b), so we have δ=2pα\delta=2p\alpha. As a consequence, we have α=β>0\alpha=\beta>0. This implies that the attracting endpoints of aa and zabz_{ab} in X\partial_{\infty}X satisfy (a(+),zab(+))<π2\sphericalangle(a({+\infty}),z_{ab}({+\infty}))<\frac{\pi}{2}.

  • \bullet

    If pp is even, then since zab=wp(a,b)z_{ab}=w_{p}(a,b), we deduce that pα+pβ=δ>0p\alpha+p\beta=\delta>0. As a consequence, α>0\alpha>0 or β>0\beta>0. This implies that (a(+),zab(+))<π2\sphericalangle(a({+\infty}),z_{ab}({+\infty}))<\frac{\pi}{2} or (b(+),zab(+))<π2\sphericalangle(b({+\infty}),z_{ab}({+\infty}))<\frac{\pi}{2}.

Proposition 4.4.

Let GG be a group acting geometrically on a CAT(0) cube complex XX, and let AA, BB be subgroups of GG which are convex-cocompact in XX, such that ABA\cap B is finite. Then for each aAa\in A, bBb\in B of infinite order, their attractive endpoints in X\partial_{\infty}X satisfy (a(+),b(+))π2\sphericalangle(a({+\infty}),b({+\infty}))\geqslant\frac{\pi}{2}.

Proof.

Let MA,MBM_{A},M_{B} denote convex cube subcomplexes of XX on which A,BA,B respectively act geometrically.

Fix xX(0)x\in X^{(0)}, and let R0R\geqslant 0 such that d1(x,MA)Rd_{1}(x,M_{A})\leqslant R and d1(x,MB)Rd_{1}(x,M_{B})\leqslant R. Let xAMAx_{A}\in M_{A} and xBMBx_{B}\in M_{B} such that d1(x,xA)Rd_{1}(x,x_{A})\leqslant R and d1(x,xB)Rd_{1}(x,x_{B})\leqslant R. Define

S={yX(0)|d1(y,MA)R and d1(y,MB)R}.S=\{y\in X^{(0)}\,|\,d_{1}(y,M_{A})\leqslant R\mbox{ and }d_{1}(y,M_{B})\leqslant R\}.

We have xSx\in S. We claim that SS is finite: if not, since XX is locally compact, we can consider a sequence (sn)n(s_{n})_{n\in\mathbb{N}} of pairwise distinct elements of SS. Since AA and BB act geometrically on MAM_{A} and MBM_{B} respectively, up to passing to a subsequence, we deduce that there exist vertices yAMAy_{A}\in M_{A}, yBMBy_{B}\in M_{B} and sequences (an)n(a_{n})_{n\in\mathbb{N}}, (bn)n(b_{n})_{n\in\mathbb{N}} of pairwise distinct elements in AA and BB respectively, such that the sequence d1(anyA,bnyB)d_{1}(a_{n}\cdot y_{A},b_{n}\cdot y_{B}) is bounded above. Since the action of GG on XX is proper and XX is locally compact, we can assume up to passing to a subsequence that the sequence (bn1an)n(b_{n}^{-1}a_{n})_{n\in\mathbb{N}} is constant, hence for all m,nm,n\in\mathbb{N} we have anam1=bnbm1ABa_{n}a_{m}^{-1}=b_{n}b_{m}^{-1}\in A\cap B. As ABA\cap B is finite, this is a contradiction. So SS is finite.

From now on, fix aAa\in A and bBb\in B of infinite order. We will show that their attractive endpoints in X\partial_{\infty}X satisfy (a(+),b(+))π2\sphericalangle(a({+\infty}),b({+\infty}))\geqslant\frac{\pi}{2}.

Let μ:X3X\mu:X^{3}\rightarrow X denote the median on XX. Fix (αn)n,(βn)n(\alpha_{n})_{n\in\mathbb{N}},(\beta_{n})_{n\in\mathbb{N}} sequences in MA(0)M_{A}^{(0)} (resp. MB(0)M_{B}^{(0)}) converging to a(+)a({+\infty}) (resp. b(+)b({+\infty})). For each nn\in\mathbb{N}, define mn=μ(αn,βn,x)m_{n}=\mu(\alpha_{n},\beta_{n},x). As αn\alpha_{n}, βn\beta_{n} and xx are vertices of XX, mnm_{n} is also a vertex of XX.

Since μ\mu is 11-Lipschitz with respect to d1d_{1} (see Lemma 1.7), we deduce that d1(mn,MA)d1(x,xA)+d1(μ(αn,βn,xA),MA)d_{1}(m_{n},M_{A})\leqslant d_{1}(x,x_{A})+d_{1}(\mu(\alpha_{n},\beta_{n},x_{A}),M_{A}). Since αn\alpha_{n} and xAx_{A} belong to the convex subcomplex MAM_{A}, we deduce that μ(αn,βn,xA)MA\mu(\alpha_{n},\beta_{n},x_{A})\in M_{A}, so d1(mn,MA)Rd_{1}(m_{n},M_{A})\leqslant R. For the same reason, we have d1(mn,MB)Rd_{1}(m_{n},M_{B})\leqslant R. As a consequence, we have n,mnS\forall n\in\mathbb{N},m_{n}\in S.

Since SS is finite, up to passing to a subsequence we may assume that n,mn=x0\forall n\in\mathbb{N},m_{n}=x_{0} is constant.

Fix ε>0\varepsilon>0, and for each nn\in\mathbb{N}, let αn\alpha^{\prime}_{n} (resp βn\beta^{\prime}_{n}) be the point on the CAT(0) geodesic segment between x0x_{0} and αn\alpha_{n} (resp. βn\beta_{n}) at d2d_{2} distance ε\varepsilon from x0x_{0} (see Figure 4).

We know that μ(x0,αn,αn)=αn\mu(x_{0},\alpha^{\prime}_{n},\alpha_{n})=\alpha^{\prime}_{n}, μ(x0,βn,βn)=βn\mu(x_{0},\beta^{\prime}_{n},\beta_{n})=\beta^{\prime}_{n} and μ(x0,αn,βn)=x0\mu(x_{0},\alpha_{n},\beta_{n})=x_{0}. Hence by using Equation (1) from Section 1.3, we deduce that

μ(x0,αn,βn)\displaystyle\mu(x_{0},\alpha_{n},\beta^{\prime}_{n}) =\displaystyle= μ(x0,αn,μ(x0,βn,βn))\displaystyle\mu(x_{0},\alpha_{n},\mu(x_{0},\beta_{n},\beta^{\prime}_{n}))
=\displaystyle= μ(μ(x0,αn,x0),μ(x0,αn,βn),βn)=μ(x0,x0,βn)=x0, and\displaystyle\mu(\mu(x_{0},\alpha_{n},x_{0}),\mu(x_{0},\alpha_{n},\beta_{n}),\beta^{\prime}_{n})=\mu(x_{0},x_{0},\beta^{\prime}_{n})=x_{0},\mbox{ and }
μ(x0,αn,βn)\displaystyle\mu(x_{0},\alpha^{\prime}_{n},\beta^{\prime}_{n}) =\displaystyle= μ(x0,μ(x0,αn,αn),βn)\displaystyle\mu(x_{0},\mu(x_{0},\alpha^{\prime}_{n},\alpha_{n}),\beta^{\prime}_{n})
=\displaystyle= μ(μ(x0,x0,βn),μ(x0,αn,βn),αn)=μ(x0,x0,αn)=x0.\displaystyle\mu(\mu(x_{0},x_{0},\beta^{\prime}_{n}),\mu(x_{0},\alpha_{n},\beta^{\prime}_{n}),\alpha^{\prime}_{n})=\mu(x_{0},x_{0},\alpha^{\prime}_{n})=x_{0}.

But the sequence (αn)n(\alpha^{\prime}_{n})_{n\in\mathbb{N}} (resp. (βn)n(\beta^{\prime}_{n})_{n\in\mathbb{N}}) actually converges to the point α\alpha^{\prime} (resp. β\beta^{\prime}) on the CAT(0) geodesic ray from x0x_{0} to a(+)a({+\infty}) (resp. b(+)b({+\infty})) at d2d_{2} distance ε\varepsilon from x0x_{0}. Hence we conclude that μ(x0,α,β)=x0\mu(x_{0},\alpha^{\prime},\beta^{\prime})=x_{0}. In other words, the path [α,x0][x0,β][\alpha^{\prime},x_{0}]\cup[x_{0},\beta^{\prime}] is monotone.

Refer to captionα\alpha^{\prime}αn\alpha^{\prime}_{n}β\beta^{\prime}βn\beta^{\prime}_{n}x0x_{0}ee
Figure 4: The proof of Proposition 4.4

On the other hand, we have x0(α,β)=x0(a(+),b(+))(a(+),b(+))\sphericalangle_{x_{0}}(\alpha^{\prime},\beta^{\prime})=\sphericalangle_{x_{0}}(a({+\infty}),b({+\infty}))\leqslant\sphericalangle(a({+\infty}),b({+\infty})). By contradiction, assume that we have (a(+),b(+))<π2\sphericalangle(a({+\infty}),b({+\infty}))<\frac{\pi}{2}, then x0(α,β)<π2\sphericalangle_{x_{0}}(\alpha^{\prime},\beta^{\prime})<\frac{\pi}{2}. Let σα,σβ\sigma_{\alpha^{\prime}},\sigma_{\beta^{\prime}} denote the minimal (closed) simplices in the link LL of XX at x0x_{0} containing α\alpha^{\prime} and β\beta^{\prime} respectively. Since x0(α,β)<π2\sphericalangle_{x_{0}}(\alpha^{\prime},\beta^{\prime})<\frac{\pi}{2}, we know that d(σα,σβ)<π2d(\sigma_{\alpha^{\prime}},\sigma_{\beta^{\prime}})<\frac{\pi}{2}.

We claim that σα\sigma_{\alpha^{\prime}} and σβ\sigma_{\beta^{\prime}} are not disjoint. Consider any two disjoint simplices σ,σ\sigma,\sigma^{\prime} in LL, we will prove that d(σ,σ)π2d(\sigma,\sigma^{\prime})\geqslant\frac{\pi}{2}. By considering a geodesic from σ\sigma to σ\sigma^{\prime} (if it exists, otherwise the distance is infinite), up to reducing the distance between σ\sigma and σ\sigma^{\prime}, we can assume that both σ\sigma and σ\sigma^{\prime} are contained in a common simplex σ0\sigma_{0}. We choose σ0\sigma_{0} minimal, so that σ0\sigma_{0} is isometric to the spherical join σσ\sigma\star\sigma^{\prime}. By definition of the spherical join, the distance between σ\sigma and σ\sigma^{\prime} is equal to π2\frac{\pi}{2}. So σα\sigma_{\alpha^{\prime}} and σβ\sigma_{\beta^{\prime}} are not disjoint.

Therefore there exists an edge ee in XX adjacent to x0x_{0} whose image in the link of x0x_{0} belongs to σασβ\sigma_{\alpha^{\prime}}\cap\sigma_{\beta^{\prime}}. So x0(α,e)<π2\sphericalangle_{x_{0}}(\alpha^{\prime},e)<\frac{\pi}{2} and x0(β,e)<π2\sphericalangle_{x_{0}}(\beta^{\prime},e)<\frac{\pi}{2}. If we consider a shifted hyperplane HH dual to ee close to x0x_{0} (the CAT(0) orthogonal of ee at a point near x0x_{0}), we see that HH separates x0x_{0} and {α,β}\{\alpha^{\prime},\beta^{\prime}\}: this contradicts the monotonicity of the path [α,x0][x0,β][\alpha^{\prime},x_{0}]\cup[x_{0},\beta^{\prime}].

As a consequence, we have (a(+),b(+))π2\sphericalangle(a({+\infty}),b({+\infty}))\geqslant\frac{\pi}{2}. ∎

We can now prove the first two results giving obstructions to being virtually cocompactly cubulated. They show how to combine Propositions 4.3 and 4.4.

Recall that a subgroup HH is said to virtually contain gg if there exists n1n\geqslant 1 such that gnHg^{n}\in H.

Lemma 4.5.

There is no group GG satisfying the following.

  • \bullet

    there exist elements a,bGa,b\in G such that a,bA(p)\langle a,b\rangle\simeq A(p), for some p3p\geqslant 3,

  • \bullet

    there exists a finite index normal subgroup G0G_{0} of GG acting geometrically on a CAT(0) cube complex XX,

  • \bullet

    there exists an abelian subgroup CC of G0G_{0} virtually containing zabz_{ab},

  • \bullet

    there exists an abelian subgroup AA of G0G_{0} virtually containing aa such that ACA\cap C is finite,

  • \bullet

    if pp is even, there exists an abelian subgroup BB of G0G_{0} virtually containing bb such that BCB\cap C is finite,

  • \bullet

    for every gGg\in G, the groups gAg1gAg^{-1} and gCg1gCg^{-1} (and gBg1gBg^{-1} if pp is even) are convex-cocompact in XX.

Proof.

By contradiction, assume that such a group GG exists. We will produce a proper action of GG on a CAT(0) space.

Consider the induced action of GG on the finite-dimensional CAT(0) cube complex XG/G0X^{G/G_{0}}. To describe this action, one can for instance identify XG/G0X^{G/G_{0}} with the space of right G0G_{0}-equivariant maps from GG to XX, endowed with the action of GG by left translations. This provides a proper action of GG on the CAT(0) cube complex XG/G0X^{G/G_{0}} by cubical isometries (this idea comes from [Bri10, Remark 1]).

Let N1N\geqslant 1 such that aNAa^{N}\in A, and let M1M\geqslant 1 such that zabMCz_{ab}^{M}\in C. We will now prove that the attractive endpoints of aNa^{N} and zabMz_{ab}^{M} in XG/G0\partial X^{G/G_{0}} satisfy XG/G0(aN(+),zabM(+))π2\sphericalangle_{\partial X^{G/G_{0}}}(a^{N}({+\infty}),z_{ab}^{M}({+\infty}))\geqslant\frac{\pi}{2}, which will contradict Proposition 4.3.

Let us denote G/G0={g1G0,,gnG0}G/G_{0}=\{g_{1}G_{0},\dots,g_{n}G_{0}\}. The action of G0G_{0} on XG/G0X^{G/G_{0}} preserves each factor, and the action of G0G_{0} on XgiG0X^{g_{i}G_{0}} is isomorphic to the conjugate by gig_{i} of the original action of G0G_{0} on XX. For each 1in1\leqslant i\leqslant n, we know that giaNgi1giAgi1g_{i}a^{N}g_{i}^{-1}\in g_{i}Ag_{i}^{-1}, and the subgroup giAgi1g_{i}Ag_{i}^{-1} is convex-cocompact in XX by assumption. Similarly, for each 1in1\leqslant i\leqslant n, we know that gizabMgi1giCgi1g_{i}z_{ab}^{M}g_{i}^{-1}\in g_{i}Cg_{i}^{-1}, and the subgroup giCgi1g_{i}Cg_{i}^{-1} is convex-cocompact in XX by assumption. Furthermore, the intersection giAgi1giCgi1=gi(AC)gi1={1}g_{i}Ag_{i}^{-1}\cap g_{i}Cg_{i}^{-1}=g_{i}(A\cap C)g_{i}^{-1}=\{1\} is finite. According to Proposition 4.4, we deduce that XgiG0(aN(+),zabM(+))π2\sphericalangle_{\partial X^{g_{i}G_{0}}}(a^{N}({+\infty}),z_{ab}^{M}({+\infty}))\geqslant\frac{\pi}{2}.

Note that the visual boundary of the finite product XG/G0X^{G/G_{0}} is isometric to the spherical join of the visual boundaries XgiG0\partial X^{g_{i}G_{0}} of each factor (see [BH99, Proposition I.5.15]). Furthermore, by definition of the distance on the spherical join S1S2S_{1}\ast S_{2} of two metric spaces S1,S2S_{1},S_{2} (see [BH99, Definition I.5.13]), if x,yS1S2x,y\in S_{1}\ast S_{2} are such that their distances in each factor are at least π2\frac{\pi}{2}, then xx and yy are at distance at least π2\frac{\pi}{2}.

Since for any 1in1\leqslant i\leqslant n we have XgiG0(aN(+),zabM(+))π2\sphericalangle_{\partial X^{g_{i}G_{0}}}(a^{N}({+\infty}),z_{ab}^{M}({+\infty}))\geqslant\frac{\pi}{2}, we deduce that XG/G0(aN(+),zabM(+))π2\sphericalangle_{\partial X^{G/G_{0}}}(a^{N}({+\infty}),z_{ab}^{M}({+\infty}))\geqslant\frac{\pi}{2}. By symmetry if pp is even, we also have XG/G0(bN(+),zabM(+))π2\sphericalangle_{\partial X^{G/G_{0}}}(b^{N}({+\infty}),z_{ab}^{M}({+\infty}))\geqslant\frac{\pi}{2}.

Remark now that the group GG contains the dihedral Artin group a,b\langle a,b\rangle, and acts properly by semisimple isometries on the CAT(0) cube complex XG/G0X^{G/G_{0}}, so this contradicts Proposition 4.3. This concludes the proof. ∎

Lemma 4.6.

There is no group GG satisfying the following.

  • \bullet

    there exist elements a,bGa,b\in G such that a,bA(p)\langle a,b\rangle\simeq A(p), for some p3p\geqslant 3,

  • \bullet

    there exists a finite index normal subgroup G0G_{0} of GG acting geometrically on a CAT(0) cube complex XX,

  • \bullet

    there exists an abelian subgroup CC of G0G_{0} commuting with aa and bb and virtually containing zabz_{ab},

  • \bullet

    there exists an abelian subgroup AA of G0G_{0} virtually containing aa such that Azab={1}A\cap\langle z_{ab}\rangle=\{1\},

  • \bullet

    if pp is even, there exists an abelian subgroup BB of G0G_{0} virtually containing bb such that ABzab={1}AB\cap\langle z_{ab}\rangle=\{1\},

  • \bullet

    for every gGg\in G, the groups gAg1gAg^{-1} and gCg1gCg^{-1} (and gBg1gBg^{-1} if pp is even) are convex-cocompact in XX.

Proof.

By contradiction, assume that there exists a counterexample GG. Assume furthermore that, among all counterexamples, the dimension of the CAT(0) cube complex XX is minimal. We will prove that GG is then a counterexample to Lemma 4.5.

We will now prove that ACA\cap C is finite. Let W=ACW=A\cap C. Since WW is abelian, according to Proposition 4.1, there exists a finite index subgroup W0WG0W_{0}\subset W\cap G_{0} such that the centralizer H=ZG(W0)H=Z_{G}(W_{0}) has a finite index normal subgroup H0G0H_{0}\subset G_{0} acting geometrically on a CAT(0) cube complex YY with dimYdimX\operatorname{dim}Y\leqslant\operatorname{dim}X. Furthermore, for every gHg\in H, the groups gAg1H0gAg^{-1}\cap H_{0} and gCg1H0gCg^{-1}\cap H_{0} (and gBg1H0gBg^{-1}\cap H_{0} if pp is even) are convex-cocompact in YY. Also note that, since a,ba,b commute with CC and WCW\subset C, we deduce that a,bHa,b\in H.

According to Proposition 4.2, the group W0W_{0} has a finite index subgroup WW^{\prime} such that G=H/WG^{\prime}=H/W^{\prime} has a finite index subgroup G0=H0/WG^{\prime}_{0}=H_{0}/W^{\prime} acting geometrically on a CAT(0) cube complex XX^{\prime}, with dimXdimYrkWdimXrkW\operatorname{dim}X^{\prime}\leqslant\operatorname{dim}Y-\operatorname{rk}W^{\prime}\leqslant\operatorname{dim}X-\operatorname{rk}W^{\prime}. Furthermore, for every gHg\in H, the groups (gAWg1H0)/W(gAW^{\prime}g^{-1}\cap H_{0})/W^{\prime} and (gCWg1H0)/W(gCW^{\prime}g^{-1}\cap H_{0})/W^{\prime} (and (gBWg1H0)/W(gBW^{\prime}g^{-1}\cap H_{0})/W^{\prime} if pp is even) are convex-cocompact in XX^{\prime}.

We will prove that the group GG^{\prime} is also a counterexample to Lemma 4.6.

  • \bullet

    We will prove that a=aWa^{\prime}=aW^{\prime} and b=bWb^{\prime}=bW^{\prime} span a subgroup of GG^{\prime} isomorphic to a,bA(p)\langle a,b\rangle\simeq A(p). Since WW^{\prime} is central in HH, we deduce that a,bW\langle a,b\rangle\cap W^{\prime} is central in a,b\langle a,b\rangle, so a,bWzabA={1}\langle a,b\rangle\cap W^{\prime}\subset\langle z_{ab}\rangle\cap A=\{1\}. Hence a,bW={1}\langle a,b\rangle\cap W^{\prime}=\{1\}, and a,ba,bA(p)\langle a^{\prime},b^{\prime}\rangle\simeq\langle a,b\rangle\simeq A(p).

  • \bullet

    The finite index subgroup G0G^{\prime}_{0} of GG^{\prime} acts geometrically on the CAT(0) cube complex XX^{\prime}.

  • \bullet

    The abelian subgroup C=(CWH)/WC^{\prime}=(CW^{\prime}\cap H)/W^{\prime} of GG^{\prime} virtually contains zabz_{a^{\prime}b^{\prime}}, since both CC and HH virtually contain zabz_{ab}. Furthermore, since aa and bb commute with CC, we deduce that aa^{\prime} and bb^{\prime} commute with CC^{\prime}.

  • \bullet

    The abelian subgroup A=(AWH)/WA^{\prime}=(AW^{\prime}\cap H)/W^{\prime} of GG^{\prime} virtually contains aa^{\prime}, since both AA and HH virtually contain aa. Furthermore, since WAW^{\prime}\subset A, we have AW(zabW)=A(zabW)=(Azab)W=WAW^{\prime}\cap(\langle z_{ab}\rangle W^{\prime})=A\cap(\langle z_{ab}\rangle W^{\prime})=(A\cap\langle z_{ab}\rangle)W^{\prime}=W^{\prime}, so Azab={1}A^{\prime}\cap\langle z_{a^{\prime}b^{\prime}}\rangle=\{1\}.

  • \bullet

    If pp is even, similarly the abelian subgroup B=(BWH)/WB^{\prime}=(BW^{\prime}\cap H)/W^{\prime} of GG^{\prime} virtually contains bb^{\prime}. Furthermore, assume that nn\in\mathbb{Z} and wWw\in W^{\prime} are such that zabnw(AWH)(BWH)z_{a^{\prime}b^{\prime}}^{n}w\in(AW^{\prime}\cap H)(BW^{\prime}\cap H), we will prove that n=0n=0. As WBWHW^{\prime}\subset BW^{\prime}\cap H, we have zabn(AWH)(BWH)z_{a^{\prime}b^{\prime}}^{n}\in(AW^{\prime}\cap H)(BW^{\prime}\cap H). Since WAW^{\prime}\subset A, we have AW=AAW^{\prime}=A so zabn(AH)(BWH)z_{a^{\prime}b^{\prime}}^{n}\in(A\cap H)(BW^{\prime}\cap H). As WBWHW^{\prime}\subset BW^{\prime}\cap H and WW^{\prime} commutes with HH, we have zabn(AH)W(BH)z_{a^{\prime}b^{\prime}}^{n}\in(A\cap H)W^{\prime}(B\cap H). Since WAHW^{\prime}\subset A\cap H, we have zabn(AH)(BH)ABz_{a^{\prime}b^{\prime}}^{n}\in(A\cap H)(B\cap H)\subset AB. By assumption, n=0n=0. So ABzab={1}A^{\prime}B^{\prime}\cap\langle z_{a^{\prime}b^{\prime}}\rangle=\{1\}.

  • \bullet

    For any gWGgW^{\prime}\in G^{\prime}, we have seen that the groups (gW)A(gW)1G0=(gAWg1H0)/W(gW^{\prime})A^{\prime}(gW^{\prime})^{-1}\cap G^{\prime}_{0}=(gAW^{\prime}g^{-1}\cap H_{0})/W^{\prime} and (gW)C(gW)1G0=(gCWg1H0)/W(gW^{\prime})C^{\prime}(gW^{\prime})^{-1}\cap G^{\prime}_{0}=(gCW^{\prime}g^{-1}\cap H_{0})/W^{\prime} (and (gW)B(gW)1G0=(gBWg1H0)/W(gW^{\prime})B^{\prime}(gW^{\prime})^{-1}\cap G^{\prime}_{0}=(gBW^{\prime}g^{-1}\cap H_{0})/W if pp is even) are convex-cocompact in XX^{\prime}.

As a consequence, the group GG^{\prime} is also a counterexample to Lemma 4.6, with dimXdimXrkW\operatorname{dim}X^{\prime}\leqslant\operatorname{dim}X-\operatorname{rk}W^{\prime}. By minimality of dimX\operatorname{dim}X, we deduce that rkW=0\operatorname{rk}W^{\prime}=0, so W=ACW=A\cap C is finite.

If pp is even, we argue similarly that BCB\cap C is finite. To be precise, the only difference in the statement between AA and BB is that Azab={1}A\cap\langle z_{ab}\rangle=\{1\} and ABzab={1}AB\cap\langle z_{ab}\rangle=\{1\}. But this assumption is, in fact, symmetrical with respect to AA and BB: indeed ABzab={1}AB\cap\langle z_{ab}\rangle=\{1\} implies both Bzab={1}B\cap\langle z_{ab}\rangle=\{1\} and BAzab={1}BA\cap\langle z_{ab}\rangle=\{1\}.

As a consequence, the group GG contradicts Lemma 4.6. Therefore there exists no such group GG. ∎

We are now ready to state the most general and self-contained result giving an obstruction to being virtually cocompactly cubulated. In the proof, we will produce small convex-cocompact subgroups using highest abelian subgroups.

Proposition 4.7.

Let GG be a group satisfying the following.

  • \bullet

    There exist elements a,bGa,b\in G such that a,bA(p)\langle a,b\rangle\simeq A(p), for some p3p\geqslant 3,

  • \bullet

    For every n1n\geqslant 1, we have ZG(an)=ZG(a)Z_{G}(a^{n})=Z_{G}(a) and ZG(bn)=ZG(b)Z_{G}(b^{n})=Z_{G}(b),

  • \bullet

    There exists αG\alpha\in G commuting with aa, such that no non-zero powers of α\alpha and zabz_{ab} commute,

  • \bullet

    If pp is even, there exists βG\beta\in G, commuting with bb, such that for every q1q\geqslant 1 we have ZG(a,zabq,αq)ZG(b,βq)zab={1}Z_{G}(a,z_{ab}^{q},\alpha^{q})Z_{G}(b,\beta^{q})\cap\langle z_{ab}\rangle=\{1\}.

Then GG is not virtually cocompactly cubulated.

Proof.

By contradiction, assume that some finite index normal subgroup G0G_{0} of GG acts geometrically on a CAT(0) cube complex XX. We will prove that GG is then a counterexample to Lemma 4.6.

Let AA denote the intersection of two highest maximal abelian subgroups of GG virtually containing {a,α}\{a,\alpha\} and {a,zab}\{a,z_{ab}\} respectively. Since for every n1n\geqslant 1 we have ZG(an)=ZG(a)Z_{G}(a^{n})=Z_{G}(a), by maximality we deduce that aAa\in A. Similarly, let BB denote a highest maximal abelian subgroup of GG virtually containing bb and β\beta: BB contains bb.

Let A,BA^{\prime},B^{\prime} denote two highest maximal abelian subgroups of GG virtually containing {a,zab}\{a,z_{ab}\} and {b,zab}\{b,z_{ab}\} respectively. As for AA and BB, we know that aAa\in A^{\prime} and bBb\in B^{\prime}. So C=ABC=A^{\prime}\cap B^{\prime} commutes with aa and bb. Furthermore CC virtually contains zabz_{ab}.

According to Theorem 3.6 and Lemma 3.4, we know that for every gGg\in G, the groups gAg1G0gAg^{-1}\cap G_{0} and gCg1G0gCg^{-1}\cap G_{0} (and gBg1G0gBg^{-1}\cap G_{0} if pp is even) are convex-cocompact in XX.

By definition of AA, there exists m1m\geqslant 1 such that αm\alpha^{m} commutes with AA. Since no non-zero powers of α\alpha and zabz_{ab} commute, we deduce that Azab={1}A\cap\langle z_{ab}\rangle=\{1\}.

If pp is even, then by definition of AA and BB we know that there exists q1q\geqslant 1 such that AZG(a,zabq,αq)A\subset Z_{G}(a,z_{ab}^{q},\alpha^{q}) and BZG(b,βq)B\subset Z_{G}(b,\beta^{q}), hence zabAB={1}\langle z_{ab}\rangle\cap AB=\{1\}.

In conclusion, GG is a counterexample to Lemma 4.6. Therefore GG is not virtually cocompactly cubulated. ∎

In order to apply Proposition 4.7 to Artin groups, we need the following technical results.

Lemma 4.8.

Assume that S={a,b,c}S=\{a,b,c\}, and M=(mst)s,tSM=(m_{st})_{s,t\in S} is a Coxeter matrix such that mabm_{ab} is finite and odd, macm_{ac} is finite and mbcm_{bc} is different from 22. Then zacz_{ac} and zabz_{ab} do not virtually commute.

Proof.

Assume there exist n,mn,m\in\mathbb{Z} such that zabnz_{ab}^{n} and zacmz_{ac}^{m} commute. We can assume that n,m0n,m\geqslant 0. Then zabnzacm=zacmzabnz_{ab}^{n}z_{ac}^{m}=z_{ac}^{m}z_{ab}^{n} is an equality between positive words, so by [Par02] they are equal in the positive monoid: one can pass from one to the other by applying only the standard relations of A(S)A(S). But the relation between bb and cc cannot be used since mbc2m_{bc}\neq 2, the subword wmbc(b,c)w_{m_{bc}}(b,c) cannot appear. As a consequence, starting from zabnzacmz_{ab}^{n}z_{ac}^{m} it is not possible to obtain a word with a letter cc on the left of a letter bb. This implies that n=0n=0 or m=0m=0: no non-trivial powers of zabz_{ab} and zacz_{ac} commute. ∎

Lemma 4.9.

Assume that S={a,b,c}S=\{a,b,c\}, and M=(mst)s,tSM=(m_{st})_{s,t\in S} is a Coxeter matrix with finite entries such that mabm_{ab} and macm_{ac} are even numbers different from 22, and mbcm_{bc} is even. Then zacz_{ac} and zabz_{ab} do not virtually commute.

Furthermore, fix q1q\geqslant 1 and assume that uA(M)u\in A(M) commutes with aa, zabqz_{ab}^{q} and zacqz_{ac}^{q}, and that vA(M)v\in A(M) commutes with bb and zbcqz_{bc}^{q}. If there exists nn\in\mathbb{Z} such that uv=zabnuv=z_{ab}^{n}, then n=0n=0.

Proof.

Assume there exist n,mn,m\in\mathbb{Z} such that zabnz_{ab}^{n} and zacmz_{ac}^{m} commute. We can assume that n,m0n,m\geqslant 0. Then zabnzacm=zacmzabnz_{ab}^{n}z_{ac}^{m}=z_{ac}^{m}z_{ab}^{n} is an equality between positive words, so by [Par02] they are equal in the positive monoid: one can pass from one to the other by applying only the standard relations of A(S)A(S). But, starting from zabnzacmz_{ab}^{n}z_{ac}^{m}, the letters c,a,bc,a,b cannot appear in that order, since mab4m_{ab}\geqslant 4 and mac4m_{ac}\geqslant 4. This implies that n=0n=0 or m=0m=0: no non-trivial powers of zabz_{ab} and zacz_{ac} commute.

Since 1mab+1mbc+1mac1\frac{1}{m_{ab}}+\frac{1}{m_{bc}}+\frac{1}{m_{ac}}\leqslant 1, the group A(M)A(M) is not of spherical type. According to [CD95b, Theorem B], the Deligne complex, with Moussong’s metric, is CAT(0). According to [God07, Theorem 1], Godelle’s Property ()(\star\star\star) can be applied to the pair ({a},{a})(\{a\},\{a\}) to conclude that the centralizer of aa in A(M)A(M) can be described, using ribbons, as

ZA(M)(a)=a,zab,zac.Z_{A(M)}(a)=\langle a,z_{ab},z_{ac}\rangle.

By Charney and Davis (see [CD95b, Theorem B] and [CD95a, Corollary 1.4.2]), the cohomological dimension of A(M)A(M) is 22, so the maximal rank of an abelian subgroup of A(M)A(M) is 22. As a consequence, the only elements of ZA(M)(a)Z_{A(M)}(a) commuting with zabqz_{ab}^{q} and zacqz_{ac}^{q} are powers of aa, so uau\in\langle a\rangle. Let rr\in\mathbb{Z} such that u=aru=a^{r}.

Assume that vA(M)v\in A(M) commutes with bb and zbcqz_{bc}^{q}, and that nn\in\mathbb{Z} is such that uv=zabnuv=z_{ab}^{n}. We will prove that n=0n=0. Remark that zbcqz_{bc}^{q} commutes with v=u1zabn=arzabnv=u^{-1}z_{ab}^{n}=a^{-r}z_{ab}^{n}.

Consider the homomorphism ϕc:A(M)a,b\phi_{c}:A(M)\rightarrow\langle a,b\rangle sending a,ba,b to a,ba,b and sending cc to 11. Since all integers defining MM are even, ϕc\phi_{c} is a well-defined group homomorphism. Then ϕc(zbcq)=bqmbc2\phi_{c}(z_{bc}^{q})=b^{\frac{qm_{bc}}{2}} commutes with ϕc(arzabn)=arzabn\phi_{c}(a^{-r}z_{ab}^{n})=a^{-r}z_{ab}^{n}, so r=0r=0.

We have u=ar=1u=a^{r}=1, so zabn=vz_{ab}^{n}=v commutes with zbcqz_{bc}^{q}, which implies that n=0n=0. ∎

Lemma 4.10.

Assume that S={a,b,c,d}S=\{a,b,c,d\}, and M=(mst)s,tSM=(m_{st})_{s,t\in S} is a Coxeter matrix with all entries even or infinite. Assume furthermore that mabm_{ab} is finite and different from 22, macm_{ac} and mbdm_{bd} are finite, and mad,mbcm_{ad},m_{bc} are different from 22. Then zacz_{ac} and zabz_{ab} do not virtually commute.

Furthermore, fix q1q\geqslant 1 and assume that uA(M)u\in A(M) commutes with aa, zabqz_{ab}^{q} and zacqz_{ac}^{q}, and that vA(M)v\in A(M) commutes with bb and zbdqz_{bd}^{q}. If there exists nn\in\mathbb{Z} such that uv=zabnuv=z_{ab}^{n}, then n=0n=0.

Proof.

Following the same proof as in Lemma 4.8, we see that since mbc2m_{bc}\neq 2, no non-zero powers of zabz_{ab} and zacz_{ac} commute.

Without loss of generality, we can assume up to passing to the corresponding quotient that mac=mbd=mcd=2m_{ac}=m_{bd}=m_{cd}=2. Since all finite entries of the Coxeter matrix are even, every irreducible spherical parabolic subgroup of A(M)A(M) has rank 11 or 22. According to [CD95b, Theorem B], the Deligne complex, with Moussong’s metric, is CAT(0). According to [God07, Theorem 1], Godelle’s Properties ()(\star) and ()(\star\star\star) can be applied to the pair ({a,c},{a,c})(\{a,c\},\{a,c\}) to conclude that the commensurator of a,c\langle a,c\rangle in A(M)A(M) is equal to a,c\langle a,c\rangle.

Since uu commutes with aa and cqc^{q}, uu commensurates a,c\langle a,c\rangle, so ua,cu\in\langle a,c\rangle. As mbc2m_{bc}\neq 2, no non-zero powers of cc and zabz_{ab} commute, so uau\in\langle a\rangle. Let rr\in\mathbb{Z} such that u=aru=a^{r}.

Assume that vA(M)v\in A(M) commutes with bb and zbdqz_{bd}^{q}, and that nn\in\mathbb{Z} is such that uv=zabnuv=z_{ab}^{n}. We will prove that n=0n=0. Remark that zbdqz_{bd}^{q} commutes with v=u1zabn=arzabnv=u^{-1}z_{ab}^{n}=a^{-r}z_{ab}^{n}.

Consider the homomorphism ϕd:A(M)a,b,c\phi_{d}:A(M)\rightarrow\langle a,b,c\rangle sending a,b,ca,b,c to a,b,ca,b,c and sending dd to 11. We deduce that ϕd(zbdq)=ϕd(bqdq)=bq\phi_{d}(z_{bd}^{q})=\phi_{d}(b^{q}d^{q})=b^{q} commutes with ϕd(arzabn)=arzabn\phi_{d}(a^{-r}z_{ab}^{n})=a^{-r}z_{ab}^{n}, so r=0r=0.

We have u=ar=1u=a^{r}=1, so zabn=vz_{ab}^{n}=v commutes with zbdqz_{bd}^{q}, which implies that n=0n=0. ∎

We will now prove that the statements of Conjecture A and Conjecture B are equivalent.

Proposition 4.11.

Let M=(mst)s,tSM=(m_{st})_{s,t\in S} be a finite Coxeter matrix. Consider the following five conditions.

  1. A.

    For each pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} is odd, either mac=mbc=m_{ac}=m_{bc}=\infty or mac=mbc=2m_{ac}=m_{bc}=2.

  2. B.

    For each distinct a,bSa,b\in S such that mabm_{ab} is even and different from 22, there is an ordering of {a,b}\{a,b\} (say a<ba<b) such that, for every cS{a,b}c\in S\char 92\relax\{a,b\}, one of the following holds:

    1. (a)

      mac=mbc=2m_{ac}=m_{bc}=2,

    2. (b)

      mac=2m_{ac}=2 and mbc=m_{bc}=\infty,

    3. (c)

      mac=mbc=m_{ac}=m_{bc}=\infty, or

    4. (d)

      macm_{ac} is even and different from 22, a<ca<c in the ordering of {a,c}\{a,c\}, and mbc=m_{bc}=\infty.

  1. 1.

    There exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} is odd, macm_{ac}\neq\infty and mbc2m_{bc}\neq 2.

  2. 2.

    There exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} and macm_{ac} are even numbers different from 22, and mbcm_{bc}\neq\infty.

  3. 3.

    There exist 44 pairwise distinct a,b,c,dSa,b,c,d\in S such that mab{2,}m_{ab}\not\in\{2,\infty\}, mac,mbdm_{ac},m_{bd}\neq\infty and mad,mbc2m_{ad},m_{bc}\neq 2.

Then A.A. and B.B. hold if and only 1.1., 2.2. and 3.3. do not hold.

Proof.

Assume first that 1.1., 2.2. or 3.3. holds, we will prove that A.A. or B.B. do not hold.

  1. 1.

    Assume that there exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} is odd, macm_{ac}\neq\infty and mbc2m_{bc}\neq 2. Then a,b,ca,b,c contradict Condition AA.

  2. 2.

    Assume that there exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} and macm_{ac} are even numbers different from 22, and mbcm_{bc}\neq\infty. Then a,b,ca,b,c contradict Condition B.(d)B.(d).

  3. 3.

    Assume that there exist 44 pairwise distinct a,b,c,dSa,b,c,d\in S such that mab{2,}m_{ab}\not\in\{2,\infty\} is even, mac,mbdm_{ac},m_{bd}\neq\infty and mad,mbc2m_{ad},m_{bc}\neq 2. If an ordering of {a,b}\{a,b\} as in Condition B.B. existed, we should have both a<ba<b and b<ab<a, which is a contradiction.

Assume now that 1.1., 2.2. and 3.3. do not hold, we will prove that A.A. and B.B. hold.

  1. A.

    Consider three pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} is odd. Since Condition 1.1. does not hold, we have mac=mbc{2,}m_{ac}=m_{bc}\in\{2,\infty\}.

  2. B.

    Consider distinct a,bSa,b\in S such that mabm_{ab} is even and different from 22. If there exists cS{a,b}c\in S\char 92\relax\{a,b\} such that macm_{ac}\neq\infty and mbc=m_{bc}=\infty, choose the ordering a<ba<b. If there exists dS{a,b}d\in S\char 92\relax\{a,b\} such that mbdm_{bd}\neq\infty and mad=m_{ad}=\infty, choose the ordering b<ab<a.If there is no such cc or dd, choose an arbitrary ordering of {a,b}\{a,b\}.

    Notice that it is not possible that both cc and dd exist. By contradiction, assume that there exist c,dS{a,b}c,d\in S\char 92\relax\{a,b\} such that mac,mbdm_{ac},m_{bd}\neq\infty and mbc,mad=m_{bc},m_{ad}=\infty. This contradicts Condition 33.

    Now that the ordering of {a,b}\{a,b\} is well-defined, say a<ba<b, we will check that it satisfies the required properties. Fix any cS{a,b}c\in S\char 92\relax\{a,b\}.

    Assume first that mac,mbc{2,}m_{ac},m_{bc}\in\{2,\infty\}. Then since a<ba<b, we do not have both mac=m_{ac}=\infty and mbc=2m_{bc}=2.

    Assume now that mbc{2,}m_{bc}\not\in\{2,\infty\}. Since Condition 1.1. does not hold, mbcm_{bc} is even. Since Condition 2.2. does not hold, we have mac=m_{ac}=\infty. This contradicts a<ba<b.

    Assume finally that mac{2,}m_{ac}\not\in\{2,\infty\}. Since Condition 1.1. does not hold, macm_{ac} is even. Since Condition 2.2. does not hold, mbc=m_{bc}=\infty. This implies that a<ca<c in the ordering of {a,c}\{a,c\}.

    As a consequence, Conditions A.A. and B.B. are satisfied.

Let us recall the definition of the property ()(\dagger) needed to prove Conjecture A. Let M=(mab)a,bSM=(m_{ab})_{a,b\in S} be a finite Coxeter matrix. We say that the Artin-Tits group A(M)A(M) satisfies property ()(\dagger) if

sS,n1,ZA(M)(sn)=ZA(M)(s).\forall s\in S,\forall n\geqslant 1,Z_{A(M)}(s^{n})=Z_{A(M)}(s).

We can now prove the following, which is a restatement of Theorem D.

Theorem 4.12.

Let M=(mab)a,bSM=(m_{ab})_{a,b\in S} be a finite Coxeter matrix such that the Artin-Tits group A(M)A(M) satisfies property ()(\dagger). Assume that at least one of the following holds

  • \bullet

    there exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} is odd, macm_{ac}\neq\infty and mbc2m_{bc}\neq 2,

  • \bullet

    there exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} and macm_{ac} are even numbers different from 22, and mbcm_{bc}\neq\infty, or

  • \bullet

    there exist 44 pairwise distinct a,b,c,dSa,b,c,d\in S such that mab{2,}m_{ab}\not\in\{2,\infty\}, mac,mbdm_{ac},m_{bd}\neq\infty and mad,mbc2m_{ad},m_{bc}\neq 2.

Then the Artin-Tits group A(M)A(M) is not virtually cocompactly cubulated.

Proof.
  • \bullet

    Assume first that there exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} is odd, macm_{ac}\neq\infty and mbc2m_{bc}\neq 2. Then by Lemma 4.8, the element zacz_{ac} commutes with aa, but zacz_{ac} and zabz_{ab} do not virtually commute. By Proposition 4.7 applied with α=zac\alpha=z_{ac}, A(M)A(M) is not virtually cocompactly cubulated.

    Assume now that this first situation does not occur. For the two remaining cases, we will apply the same strategy.

  • \bullet

    Assume that there exist 33 pairwise distinct a,b,cSa,b,c\in S such that mabm_{ab} and macm_{ac} are even numbers different from 22, and mbcm_{bc} is finite and even. Let α=zac\alpha=z_{ac} and β=zbc\beta=z_{bc}. According to Lemma 4.9, α\alpha and aa commute, but no non-zero powers of α\alpha and zabz_{ab} commute.

    Fix q1q\geqslant 1, and assume that uZA(M)(a,zabq,αq)u\in Z_{A(M)}(a,z_{ab}^{q},\alpha^{q}), vzA(M)(b,βq)v\in z_{A(M)}(b,\beta^{q}) and nn\in\mathbb{Z} are such that uv=zabnuv=z_{ab}^{n}. We will prove that n=0n=0.

    Consider the homomorphism ϕ:A(M)a,b,c\phi:A(M)\rightarrow\langle a,b,c\rangle sending a,b,ca,b,c to a,b,ca,b,c and sending every element in S{a,b,c}S\char 92\relax\{a,b,c\} to 11. According to the first case of the proof, we can assume that for every dS{a,b,c}d\in S\char 92\relax\{a,b,c\}, the exponents mad,mbdm_{ad},m_{bd} and mcdm_{cd} are even or infinite. Hence ϕ\phi is a well-defined group homomorphism. Hence ϕ(u)ϕ(v)=ϕ(zabn)=zabn\phi(u)\phi(v)=\phi(z_{ab}^{n})=z_{ab}^{n}.

    Then ϕ(u)\phi(u) commutes with ϕ(a)=a\phi(a)=a, ϕ(zabq)=zabq\phi(z_{ab}^{q})=z_{ab}^{q} and ϕ(αq)=ϕ(zacq)=zacq\phi(\alpha^{q})=\phi(z_{ac}^{q})=z_{ac}^{q}, and ϕ(v)\phi(v) commutes with ϕ(b)=b\phi(b)=b and ϕ(βq)=ϕ(zbcq)=zbcq\phi(\beta^{q})=\phi(z_{bc}^{q})=z_{bc}^{q}. According to Lemma 4.9 applied to ϕ(u)\phi(u) and ϕ(v)\phi(v), we have n=0n=0.

    According to Proposition 4.7, the group A(M)A(M) is not virtually cocompactly cubulated.

  • \bullet

    Assume that there exist 44 pairwise distinct a,b,c,dSa,b,c,d\in S such that mab{2,}m_{ab}\not\in\{2,\infty\}, mac,mbdm_{ac},m_{bd}\neq\infty and mad,mbc2m_{ad},m_{bc}\neq 2. Let α=zac\alpha=z_{ac} and β=zbd\beta=z_{bd}. According to Lemma 4.10, α\alpha and aa commute, but no non-zero powers of α\alpha and zabz_{ab} commute.

    Fix q1q\geqslant 1, and assume that uZA(M)(a,zabq,αq)u\in Z_{A(M)}(a,z_{ab}^{q},\alpha^{q}), vzA(M)(b,βq)v\in z_{A(M)}(b,\beta^{q}) and nn\in\mathbb{Z} are such that uv=zabnuv=z_{ab}^{n}. We will prove that n=0n=0.

    Consider the homomorphism ϕ:A(M)a,b,c,d\phi:A(M)\rightarrow\langle a,b,c,d\rangle sending a,b,c,da,b,c,d to a,b,c,da,b,c,d and sending every element in S{a,b,c,d}S\char 92\relax\{a,b,c,d\} to 11. According to the first case of the proof, we can assume that for every tS{a,b,c,d}t\in S\char 92\relax\{a,b,c,d\}, the exponents mat,mbt,mctm_{at},m_{bt},m_{ct} and mdtm_{dt} are even or infinite. Hence ϕ\phi is a well-defined group homomorphism. Hence ϕ(u)ϕ(v)=ϕ(zabn)=zabn\phi(u)\phi(v)=\phi(z_{ab}^{n})=z_{ab}^{n}.

    Then ϕ(u)\phi(u) commutes with ϕ(a)=a\phi(a)=a, ϕ(zabq)=zabq\phi(z_{ab}^{q})=z_{ab}^{q} and ϕ(αq)=ϕ(zacq)=zacq\phi(\alpha^{q})=\phi(z_{ac}^{q})=z_{ac}^{q}, and ϕ(v)\phi(v) commutes with ϕ(b)=b\phi(b)=b and ϕ(βq)=ϕ(zbdq)=zbdq\phi(\beta^{q})=\phi(z_{bd}^{q})=z_{bd}^{q}. According to Lemma 4.10 applied to ϕ(u)\phi(u) and ϕ(v)\phi(v), we have n=0n=0.

    According to Proposition 4.7, the group A(M)A(M) is not virtually cocompactly cubulated.

5 Cubulation of Artin groups

5.1 Cubulation of dihedral Artin groups

Brady and McCammond showed (see [BM00]) that for all p{2,,}p\in\{2,\ldots,\infty\}, the dihedral Artin group A(p)A(p) is cocompactly cubulated. Let us recall their construction, which will be useful. We will need this construction when p{2,}p\not\in\{2,\infty\}, but it works as well when p=2p=2, so let us fix pp\neq\infty (when p=p=\infty, the Artin group is just the rank 22 free group).

The Artin group A(p)A(p) has the following presentation, due to Brady and McCammond:

A(p)=x,a1,,ap|1ip,aiai+1=x,A(p)=\langle x,a_{1},\ldots,a_{p}\,|\,\forall 1\leqslant i\leqslant p,a_{i}a_{i+1}=x\rangle,

where ap+1=a1a_{p+1}=a_{1}. This can easily be seen, with a1a_{1} and a2a_{2} corresponding to the standard generators of A(p)A(p). The presentation 22-complex KK is a K(π,1)K(\pi,1) for A(p)A(p) consisting of 11 vertex vv, p+1p+1 loops a1,,ap,xa_{1},\ldots,a_{p},x and pp triangles a1a2x1,,apa1x1a_{1}a_{2}x^{-1},\ldots,a_{p}a_{1}x^{-1} (see Figure 5).

Refer to captionxxxxxxxxxxxxxxa1a_{1}a2a_{2}a3a_{3}apa_{p}ap+1=a1a_{p+1}=a_{1}
Figure 5: Brady and McCammond’s presentation 22-complex KK

We will define another K(π,1)K(\pi,1) for A(p)A(p), which will be cubical and will have the same underlying topological space as KK. Start with two vertices vv and ww, and p+2p+2 oriented edges between vv and ww labelled α1,,αp,β1,γ\alpha_{1},\ldots,\alpha_{p},\beta^{-1},\gamma. Finally, add the pp squares with boundary labeled by α1β1α2γ1,,αpβ1α1γ1\alpha_{1}\beta^{-1}\alpha_{2}\gamma^{-1},\ldots,\alpha_{p}\beta^{-1}\alpha_{1}\gamma^{-1} and let X(A(p))X(A(p)) denote the resulting cube complex. It is easy to see that the underlying topological space of X(A(p))X(A(p)) is homoeomorphic to KK: ww corresponds to the midpoint of the edge xx, the edge xx corresponds to the path γβ1\gamma\beta^{-1}, and each square corresponds to the union of the halves of two triangles of KK (see Figure 6).

Refer to captionγ\gammaα1\alpha_{1}β\betaγ\gammaβ\betaγ\gammaβ\betaγ\gammaβ\betaγ\gammaβ\betaγ\gammaβ\betaγ\gammaβ\betaα2\alpha_{2}α3\alpha_{3}αp\alpha_{p}
Figure 6: The square complex X(A(p))X(A(p))

Hence X(A(p))X(A(p)) is also a K(π,1)K(\pi,1) for A(p)A(p). Furthermore, one has a complete description of the link of the vertex vv: it has p+2p+2 vertices labeled α1,,αp,β1,γ\alpha_{1},\ldots,\alpha_{p},\beta^{-1},\gamma, and it is the complete bipartite graph on {α1,,αp}\{\alpha_{1},\dots,\alpha_{p}\} and {β1,γ}\{\beta^{-1},\gamma\}. This graph has no double edges and no triangle, so it is a flag simplicial complex. The description of the link of ww is similar, with all orientations reversed, so we deduce that X(A(p))X(A(p)) is a locally CAT(0) square complex. In particular, A(p)A(p) is cocompactly cubulated.

Remark.

Notice that X(A(p))X(A(p)) is naturally isometric to the product of \mathbb{R} and the infinite pp-regular tree. In the case of the 33-strand braid group B3A(3)B_{3}\simeq A(3), one recovers in the central quotient the action of B3/Z(B3)PSL(2,)/2/3B_{3}/Z(B_{3})\simeq\operatorname{PSL}(2,\mathbb{Z})\simeq\mathbb{Z}/2\mathbb{Z}*\mathbb{Z}/3\mathbb{Z} on its Bass-Serre 33-regular tree.

5.2 Recubulation of even dihedral Artin groups

In the case where pp is even, there are two other natural CAT(0) square complexes on which the dihedral Artin group A(p)A(p) acts geometrically. Each will be associated with one of the two generators aa,bb of A(p)A(p). We will describe the first one, associated with a=a1a=a_{1}.

Start with the same presentation 22-complex KK as before, and remove all edges a2,a4,,apa_{2},a_{4},\dots,a_{p} with even labels, and replace each pair of triangles (a2i+1a2i+2x1,a2i+2a2i+3x1)(a_{2i+1}a_{2i+2}x^{-1},a_{2i+2}a_{2i+3}x^{-1}), for 0ip210\leqslant i\leqslant\frac{p}{2}-1, by a square with edges a2i+1xa2i+31x1a_{2i+1}xa_{2i+3}^{-1}x^{-1}. We obtain a square complex Xa(A(p))X_{a}(A(p)) with one vertex vv, p2+1\frac{p}{2}+1 edges x,a1,a3,,ap1x,a_{1},a_{3},\dots,a_{p-1} and p2\frac{p}{2} squares a1xa31x1,,ap1xa11x1a_{1}x{a_{3}}^{-1}x^{-1},\dots,a_{p-1}x{a_{1}}^{-1}x^{-1} (see Figure 7).

Refer to captiona1a_{1}a3a_{3}ap+1=a1a_{p+1}=a_{1}xxxxxxxxxxxx
Figure 7: The square complex Xa(A(p))X_{a}(A(p))

The underlying topological space of Xa(A(p))X_{a}(A(p)) is KK, so it is also a K(π,1)K(\pi,1) for A(p)A(p). Furthermore, one has a complete description of the link of the vertex vv: it has p+2p+2 vertices labeled x,x1,a1,a11,a3,a31,,ap1,ap11x,x^{-1},a_{1},a_{1}^{-1},a_{3},a_{3}^{-1},\dots,a_{p-1},a_{p-1}^{-1}, and it is the complete bipartite graph on {x,x1}\{x,x^{-1}\} and {a1,a11,a3,a31,,ap1,ap11}\{a_{1},a_{1}^{-1},a_{3},a_{3}^{-1},\dots,a_{p-1},a_{p-1}^{-1}\}. This graph has no double edges and no triangle, so it is a flag simplicial complex. So we deduce that Xa(A(p))X_{a}(A(p)) is a locally CAT(0) square complex.

The other locally CAT(0) square complex, denoted Xb(A(p))X_{b}(A(p)), is obtained by keeping only the edges with even labels and removing those with odd labels.

The fundamental difference of Xa(A(p))X_{a}(A(p)) and X(A(p))X(A(p)) is that, in the universal covers, the visual angles between the attractive fixed points of aa and zabz_{ab} differ: in X(A(p))X(A(p)) that angle is acute, while in Xa(A(p))X_{a}(A(p)) it is equal to π2\frac{\pi}{2}. This is due to the fact that, in Xa(A(p))X_{a}(A(p)), the edge a=a1a=a_{1} belongs to the complex, so the subgroup a\langle a\rangle is convex-cocompact in Xa(A(p))X_{a}(A(p)) but not in X(A(p))X(A(p)). This illustrates the case where pp is even in Proposition 4.3.

In view of Lemma 4.5, one can see that it is not possible to find a CAT(0) cube complex with a geometric action of A(p)A(p) where both a\langle a\rangle and b\langle b\rangle are convex-cocompact. And if pp is odd, it is not even possible to find one where either a\langle a\rangle or b\langle b\rangle is convex-cocompact.

5.3 Cubulation of Artin groups of even stars

Let M=(mab)a,bSM=(m_{ab})_{a,b\in S} be a finite Coxeter matrix, which is an “even star”: there exists a “central vertex” aSa\in S such that b,cS{a},mbc=\forall b,c\in S\char 92\relax\{a\},m_{bc}=\infty and bS{a},mab\forall b\in S\char 92\relax\{a\},m_{ab} is even.

We will now prove a particular case of Theorem C, namely showing that A(M)A(M) is cocompactly cubulated. Note that J. Huang, K. Jankiewicz and P. Przytycki independently gave the same construction in [HJP16].

Write S{a}={b1,,bm}S\char 92\relax\{a\}=\{b_{1},\dots,b_{m}\}. For each 1im1\leqslant i\leqslant m, the subgroup A({abi})A(\{ab_{i}\}) of A(M)A(M) spanned by aa and bib_{i} is a dihedral Artin group with even integer: let Xa(A({abi}))X_{a}(A(\{ab_{i}\})) denote the previously constructed locally CAT(0) square complex with fundamental group A({abi})A(\{ab_{i}\}), where some edge eie_{i} in Xa(A({abi}))X_{a}(A(\{ab_{i}\})) represents aa.

Consider now the square complex X(A(M))X(A(M)) which is the gluing of the square complexes Xa(A({ab1})),,Xa(A({abm}))X_{a}(A(\{ab_{1}\})),\dots,X_{a}(A(\{ab_{m}\})) where all edges e1,,eme_{1},\dots,e_{m} are identified with a single edge ee. By Van Kampen Theorem, the fundamental group of X(A(M))X(A(M)) is the free product of A({ab1}),,A({abm})A(\{ab_{1}\}),\dots,A(\{ab_{m}\}) amalgamated over the cyclic subgroup a\langle a\rangle, which is precisely isomorphic to the Artin group A(M)A(M).

If two squares Q,QQ,Q^{\prime} in X(A(M))X(A(M)) which do not lie in the same Xa(A({abi}))X_{a}(A(\{ab_{i}\})) share an edge, this edge is ee. As a consequence, for any triple of square Q,Q,Q′′Q,Q^{\prime},Q^{\prime\prime} of X(A(M))X(A(M)) which do not lie in the same Xa(A({abi}))X_{a}(A(\{ab_{i}\})), if they pairwise share an edge, then their triple intersection is ee. So X(A(M))X(A(M)) is a locally CAT(0) square complex. As a consequence, A(M)A(M) is cocompactly cubulated.

5.4 General case

Let M=(mab)a,bSM=(m_{ab})_{a,b\in S} be a finite Coxeter matrix satisfying the two conditions of Conjecture A. We will show that the Artin group A(M)A(M) is cocompactly cubulated.

Let S0={aS|bS{a},mab{2,}}S_{0}=\{a\in S\,|\,\forall b\in S\char 92\relax\{a\},m_{ab}\in\{2,\infty\}\} denote the set of vertices having all incident labels equal to 22 or \infty.

Let S1={a1,b1},,Sn={an,bn}S_{1}=\{a_{1},b_{1}\},\ldots,S_{n}=\{a_{n},b_{n}\} denote the pairs of vertices of SS for which the edge aibia_{i}b_{i} has an odd label, for 1in1\leqslant i\leqslant n (possibly n=0n=0).

Let Sn+1,,Sn+pS_{n+1},\dots,S_{n+p} denote the subsets of vertices of SS0S\char 92\relax S_{0} for which the induced matrix MSn+i×Sn+iM_{S_{n+i}\times S_{n+i}} is an even star with central vertex aiSn+ia_{i}\in S_{n+i}, for 1ip1\leqslant i\leqslant p (possibly p=0p=0).

By assumption, we have S=0in+pSiS=\bigsqcup_{0\leqslant i\leqslant n+p}S_{i}.


We will consider cube complexes with edges labeled in 𝒫(S){\cal P}(S), the power set of SS.

Let X0X_{0} be the Salvetti cube complex of the right-angled Artin group of the graph induced by S0S_{0}: we will recall here its construction (see [Sal87]). It has one vertex and its edge set is S0S_{0}: each edge is labeled by {a}\{a\}, for some aS0a\in S_{0}. For each simplex TS0T\subset S_{0}, we add a |T||T|-cube, by identifying each of the |T||T| parallel classes of edges of [0,1]|T|[0,1]^{|T|} with the edges labeled by {t}\{t\}, for tTt\in T. Then by Theorem 1.2, X0X_{0} is locally CAT(0) cube complex.

For each 1in1\leqslant i\leqslant n, let XiX_{i} denote a copy of the previously constructed cube complex X(A(pi))X(A(p_{i})) for the subgroup generated by aia_{i} and bib_{i}, where pip_{i} is odd. Label each edge of XiX_{i} by {ai,bi}\{a_{i},b_{i}\}.

For each n+1in+pn+1\leqslant i\leqslant n+p, let XiX_{i} denote a copy of the previously constructed cube complex X(A(Si))X(A(S_{i})) for the subgroup generated by SiS_{i}. Label the edge corresponding to the element aia_{i} by {ai}\{a_{i}\}, and label each other edge coming from the square complex Xai(A({aib}))X_{a_{i}}(A(\{a_{i}b\})) by {ai,b}\{a_{i},b\}, for every bSi{ai}b\in S_{i}\char 92\relax\{a_{i}\}.

Consider the following cube complex XX, which will be a cube subcomplex of the direct product i=0n+pXi\prod_{i=0}^{n+p}X_{i}. For each set of cubes Q0,,Qn+pQ_{0},\dots,Q_{n+p} of X0,,Xn+pX_{0},\dots,X_{n+p} respectively, we will add the cube Q0××Qn+pQ_{0}\times\dots\times Q_{n+p} to XX if and only if the following holds:

0ijn+p,for any bi belonging to the label of some edge of Qi,\displaystyle\forall 0\leqslant i\neq j\leqslant n+p,\mbox{for any $b_{i}$ belonging to the label of some edge of $Q_{i}$},
for any bj belonging to the label of some edge of Qjbi and bj commute.\displaystyle\mbox{for any $b_{j}$ belonging to the label of some edge of $Q_{j}$, $b_{i}$ and $b_{j}$ commute}.

We can now give a proof of Theorem C, which we restate here.

Theorem 5.1.

XX is a locally CAT(0) cube complex, so A(M)A(M) is cocompactly cubulated.

Proof.

Let QQ,QQ^{\prime}, Q′′Q^{\prime\prime} be cubes of XX, which pairwise intersect in codimension 11, and intersect globally in codimension 22. Write Q=i=0n+pQiQ=\prod_{i=0}^{n+p}Q_{i}, Q=i=0n+pQiQ^{\prime}=\prod_{i=0}^{n+p}Q^{\prime}_{i} and Q′′=i=0n+pQi′′Q^{\prime\prime}=\prod_{i=0}^{n+p}Q^{\prime\prime}_{i}.

Let k,k,k′′0,n+pk,k^{\prime},k^{\prime\prime}\in\llbracket 0,n+p\rrbracket be such that

ik′′,Qi=Qi,ik,Qi=Qi′′ and ik,Qi=Qi′′.\forall i\neq k^{\prime\prime},Q_{i}=Q^{\prime}_{i},\forall i\neq k^{\prime},Q_{i}=Q^{\prime\prime}_{i}\mbox{ and }\forall i\neq k,Q^{\prime}_{i}=Q^{\prime\prime}_{i}.
  • \bullet

    Assume first that k=k=k′′k=k^{\prime}=k^{\prime\prime}. Then the three cubes QkQ_{k},QkQ^{\prime}_{k} and Qk′′Q^{\prime\prime}_{k} of XkX_{k} pairwise intersect in codimension 11 and globally intersect in codimension 22. Since XkX_{k} is locally CAT(0), there exists a cube KkK_{k} in XkX_{k} such that QkQ_{k},QkQ^{\prime}_{k} and Qk′′Q^{\prime\prime}_{k} are codimension 11 faces of KkK_{k}. Since for every 1in+p1\leqslant i\leqslant n+p, XiX_{i} is a square complex and KkK_{k} has dimension at least 33, we deduce that k=0k=0.

    Let K=K0×i=1n+pKiK=K_{0}\times\prod_{i=1}^{n+p}K_{i}, where 1in+p,Ki=Qi=Qi=Qi′′\forall 1\leqslant i\leqslant n+p,K_{i}=Q_{i}=Q^{\prime}_{i}=Q^{\prime\prime}_{i}.

    We will check that the cube KK belongs to XX: fix 0ijn+p0\leqslant i\neq j\leqslant n+p, and choose bib_{i} belonging to the label of some edge of KiK_{i} and bjb_{j} belonging to the label of some edge of KjK_{j}.

    • \star

      If i,j0i,j\neq 0, then Ki=QiK_{i}=Q_{i} and Kj=QjK_{j}=Q_{j}, and since QQ is a cube of XX, bib_{i} and bjb_{j} commute.

    • \star

      If i=0i=0 or j=0j=0, assume that i=0i=0. Then some edge of K0K_{0} has label {b0}\{b_{0}\}. By definition of X0X_{0}, parallel edges in K0K_{0} have the same labels, so {b0}\{b_{0}\} is also the label of some edge of Q0Q_{0}, Q0Q^{\prime}_{0} or Q0′′Q^{\prime\prime}_{0}: assume that {b0}\{b_{0}\} is the label of some edge of Q0Q_{0}. Since QQ is a cube of XX, b0b_{0} and bjb_{j} commute.

    As a consequence, KK is a cube of XX.

  • \bullet

    Assume now that k,k,k′′k,k^{\prime},k^{\prime\prime} are not all equal. Then k,k,k′′k,k^{\prime},k^{\prime\prime} are pairwise distinct. Let K=K0×i=1n+pKiK=K_{0}\times\prod_{i=1}^{n+p}K_{i}, where for each i{k,k,k′′}i\not\in\{k,k^{\prime},k^{\prime\prime}\} we have Ki=Qi=Qi=Qi′′K_{i}=Q_{i}=Q^{\prime}_{i}=Q^{\prime\prime}_{i}, and also Kk=Qk=Qk′′K_{k}=Q^{\prime}_{k}=Q^{\prime\prime}_{k}, Kk=Qk=Qk′′K_{k^{\prime}}=Q_{k^{\prime}}=Q^{\prime\prime}_{k^{\prime}} and Kk′′=Qk′′=Qk′′K_{k^{\prime\prime}}=Q_{k^{\prime\prime}}=Q^{\prime}_{k^{\prime\prime}}. Note that QkQ_{k} has codimension 11 in Qk=KkQ^{\prime}_{k}=K_{k}, and similarly QkQ^{\prime}_{k^{\prime}} has codimension 11 in KkK_{k^{\prime}} and Qk′′′′Q^{\prime\prime}_{k^{\prime\prime}} has codimension 11 in Kk′′K_{k^{\prime\prime}}. Hence KK is a cube of the product i=0n+pXi\prod_{i=0}^{n+p}X_{i}, which contains each of Q,Q,Q′′Q,Q^{\prime},Q^{\prime\prime} with codimension 11.

    We will now check that KK is a cube of XX. For any 0ijn+p0\leqslant i\neq j\leqslant n+p, assume that bib_{i} belongs to the label of some edge of KiK_{i}, and that bjb_{j} belongs to the label of some edge of KjK_{j}. Without loss of generality, ii and jj are both different from kk. So Ki=QiK_{i}=Q_{i} and Kj=QjK_{j}=Q_{j}, and since QQ is a cube of XX we know that bib_{i} and bjb_{j} commute. Hence KK is a cube of XX.

According to Theorem 1.2, XX is a locally CAT(0) cube complex.

The fundamental group of XX is given by its 22-skeleton, and it is the quotient of the free product of A(Γ|S0),,A(Γ|Sn+p)A(\Gamma|_{S_{0}}),\dots,A(\Gamma|_{S_{n+p}}) obtained by adding the following commutation relations:

0ijn+p,if aiSi and ajSj commute in A(M),ai and aj commute in π1(X).\forall 0\leqslant i\neq j\leqslant n+p,\mbox{if $a_{i}\in S_{i}$ and $a_{j}\in S_{j}$ commute in $A(M)$},\mbox{$a_{i}$ and $a_{j}$ commute in $\pi_{1}(X)$}.

The group π1(X)\pi_{1}(X) is therefore isomorphic to A(M)A(M).

As a consequence, A(M)A(M) is cocompactly cubulated. ∎

We can now give the proof of Theorem E, which we restate here.

Theorem E.

Conjecture A holds for any Artin-Tits group satisfying property ()(\dagger). In particular, Conjecture A holds for Artin-Tits groups of type FC, and for Artin-Tits groups whose irreducible spherical parabolic subgroups have rank at most 22.

Proof.

Theorem C and Theorem D precisely state that Conjecture A holds for any Artin-Tits group satisfying property ()(\dagger).

According to [God07, Theorem 1], if the Deligne complex of an Artin-Tits group can be endowed with a piecewise Euclidean CAT(0) metric, then this Artin-Tits group satisfies property ()(\dagger).

Consider an Artin-Tits group A(M)A(M) of type FC. According to [CD95b], the Deligne complex of A(M)A(M), endowed with the cubical metric, is CAT(0). Therefore A(M)A(M) satisfies property ()(\dagger), and satisfies Conjecture A.

Consider an Artin-Tits group A(M)A(M) whose irreducible spherical parabolic subgroups have rank at most 22. According to [CD95b], the Deligne complex of A(M)A(M), endowed with the Moussong metric, is CAT(0). Therefore A(M)A(M) satisfies property ()(\dagger), and satisfies Conjecture A. ∎

We can now prove Corollary H, which we restate here.

Corollary H.

The nn-strand braid group BnB_{n}, or its central quotient Bn/Z(Bn)B_{n}/Z(B_{n}), is virtually cocompactly cubulated if and only if n4n\leqslant 4.

Proof.

We have seen that B3B_{3} and B3/Z(B3)B_{3}/Z(B_{3}) are cocompactly cubulated.

Assume that n4n\geqslant 4, and let σ1,,σn1\sigma_{1},\dots,\sigma_{n-1} denote the standard generators of BnB_{n}. Since mσ1,σ2=mσ2,σ3=3m_{\sigma_{1},\sigma_{2}}=m_{\sigma_{2},\sigma_{3}}=3, according to Theorem D, BnB_{n} is not virtually cocompactly cubulated.

We will show that if the central quotient Bn/Z(Bn)B_{n}/Z(B_{n}) were virtually cocompactly cubulated, then BnB_{n} itself would be virtually cocompactly cubulated. Assume that G=Bn/Z(Bn)G=B_{n}/Z(B_{n}) has a finite index subgroup G0G_{0} acting geometrically on a CAT(0) cube complex XX. Then H=G0Z(Bn)H=G_{0}Z(B_{n}) is a finite index subgroup of BnB_{n} acting cocompatly on XX. Consider the counting morphism BnB_{n}\rightarrow\mathbb{Z} sending each σi\sigma_{i} to 11. If we compose this morphism with the standard action of \mathbb{Z} on \mathbb{R} (seen as a cube complex) by translation, this defines a cocompact action of HH on \mathbb{R}. Since the centre Z(Bn)Z(B_{n}) acts properly on \mathbb{R}, we deduce that the action of HH on X×X\times\mathbb{R} is proper and cocompact. So BnB_{n} would be virtually cocompactly cubulated.

As a consequence, Bn/Z(Bn)B_{n}/Z(B_{n}) is not virtually cocompactly cubulated. ∎

We now give a proof that Conjecture B implies Conjecture F.

Proof.

Assume that an Artin-Tits group A(M)A(M) is not of type FC, we want to prove that at least one of the three cases described in Conjecture B occur. Since A(M)A(M) is not of type FC, there exist three pairwise distinct a,b,cSa,b,c\in S such that mab,mbc,macm_{ab},m_{bc},m_{ac} are finite and 1mab+1mbc+1mac1\frac{1}{m_{ab}}+\frac{1}{m_{bc}}+\frac{1}{m_{ac}}\leqslant 1.

Assume first that at least one of mab,mbc,macm_{ab},m_{bc},m_{ac} is odd, for instance mabm_{ab} is odd. Then at least one of mbc,macm_{bc},m_{ac} is strictly bigger than 22, for instance mbc>2m_{bc}>2. Hence a,b,ca,b,c correspond to the first case described in Conjecture B.

Assume now that all of mab,mbc,macm_{ab},m_{bc},m_{ac} are even. Then at most one of mab,mbc,macm_{ab},m_{bc},m_{ac} is equal to 22, for instance mab4m_{ab}\geqslant 4 and mac4m_{ac}\geqslant 4. So a,b,ca,b,c correspond to the second case described in Conjecture B. ∎

To conclude, we give a proof of Corollary G.

Proof.

Assume first that AA is an Artin-Tits group of spherical type such that all its irreducible parabolic subgroups have rank at most 22. Then AA is a direct product of cyclic groups and dihedral Artin groups, and according to Theorem C (or according to [BM00]), AA is cocompactly cubulated.

Assume now that AA is an Artin-Tits group of spherical type containing an irreducible parabolic subgroup of rank at least 33. So there exist three pairwise distinct a,b,cSa,b,c\in S such that mab=3m_{ab}=3, macm_{ac}\neq\infty and mbc2m_{bc}\neq 2. The elements a,b,ca,b,c correspond to the first case in Conjecture B, so according to Theorem E, AA is not virtually cocompactly cubulated. ∎

References

  • [BH83] H.-J. Bandelt & J. Hedlíková – “Median algebras”, Discrete Math. 45 (1983), no. 1, p. 1–30.
  • [BH99] M. R. Bridson & A. Haefliger – Grund. math. Wiss., Metric  spaces  of  non-positive  curvature 319​ , Springer, 1999.
  • [BHS14] J. Behrstock, M. Hagen & A. Sisto – “Hierarchically hyperbolic spaces I: curve complexes for cubical groups”, (2014), arXiv:1412.2171.
  • [BM00] T. Brady & J. P. McCammond – “Three-generator Artin groups of large type are biautomatic”, J. Pure Appl. Algebra 151 (2000), no. 1, p. 1–9.
  • [BM10] T. Brady & J. McCammond – “Braids, posets and orthoschemes”, Algebr. Geom. Topol. 10 (2010), no. 4, p. 2277–2314.
  • [Bow13] B. H. Bowditch – “Coarse median spaces and groups”, Pacific J. Math. 261 (2013), no. 1, p. 53–93.
  • [Bra01] T. Brady – “A partial order on the symmetric group and new K(π,1)K(\pi,1)’s for the braid groups”, Adv. Math. 161 (2001), no. 1, p. 20–40.
  • [Bri10] M. R. Bridson – “Semisimple actions of mapping class groups on CAT(0){\rm CAT}(0) spaces”, London Math. Soc. Lecture Note Ser., in Geometry of Riemann surfaces 368 , Cambridge Univ. Press, Cambridge, 2010, p. 1–14.
  • [BS72] E. Brieskorn & K. Saito – “Artin-Gruppen und Coxeter-Gruppen”, Invent. Math. 17 (1972), p. 245–271.
  • [CD95a] R. Charney & M. W. Davis – “Finite K(π,1)K(\pi,1)s for Artin groups”, Ann. of Math. Stud., in Prospects in topology (Princeton, NJ, 1994) 138 , Princeton Univ. Press, Princeton, NJ, 1995, p. 110–124.
  • [CD95b] — , “The K(π,1)K(\pi,1)-problem for hyperplane complements associated to infinite reflection groups”, J. Amer. Math. Soc. 8 (1995), no. 3, p. 597–627.
  • [Cha] R. Charney – “Problems related to Artin groups”, (), American Institute of Mathematics, http://people.brandeis.edu/~charney/papers/_probs.pdf.
  • [Che00] V. Chepoi – “Graphs of some CAT(0){\rm CAT}(0) complexes”, Adv. in Appl. Math. 24 (2000), no. 2, p. 125–179.
  • [CM05] P.-E. Caprace & B. Mühlherr – “Reflection triangles in Coxeter groups and biautomaticity”, J. Group Theory 8 (2005), no. 4, p. 467–489.
  • [CN05] I. Chatterji & G. Niblo – “From wall spaces to CAT(0)\rm CAT(0) cube complexes”, Internat. J. Algebra Comput. 15 (2005), no. 5-6, p. 875–885.
  • [CP05] J. Crisp & L. Paoluzzi – “On the classification of CAT(0) structures for the 4-string braid group”, Michigan Math. J. 53 (2005), no. 1, p. 133–163.
  • [Del72] P. Deligne – “Les immeubles des groupes de tresses généralisés”, Invent. Math. 17 (1972), p. 273–302.
  • [God07] E. Godelle – “Artin-Tits groups with CAT(0) Deligne complex”, J. Pure Appl. Algebra 208 (2007), no. 1, p. 39–52.
  • [Gro87] M. Gromov – “Hyperbolic groups”, Math. Sci. Res. Inst. Publ., in Essays in group theory , Springer, New York, 1987, p. 75–263.
  • [Hag07] F. Haglund – “Isometries of CAT(0) cube complexes are semi-simple”, (2007), arXiv:0705.3386.
  • [HJP16] J. Huang, K. Jankiewicz & P. Przytycki – “Cocompactly cubulated 2-dimensional Artin groups”, Comment. Math. Helv. 91 (2016), no. 3, p. 519–542.
  • [HKS16] T. Haettel, D. Kielak & P. Schwer – “The 6-strand braid group is CAT(0){\rm CAT}(0)”, Geom. Dedicata 182 (2016), p. 263–286.
  • [Nic04] B. Nica – “Cubulating spaces with walls”, Algebr. Geom. Topol. 4 (2004), p. 297–309 (electronic).
  • [NR03] G. A. Niblo & L. D. Reeves – “Coxeter groups act on CAT(0){\rm CAT}(0) cube complexes”, J. Group Theory 6 (2003), no. 3, p. 399–413.
  • [Par02] L. Paris – “Artin monoids inject in their groups”, Comment. Math. Helv. 77 (2002), no. 3, p. 609–637.
  • [Sag95] M. Sageev – “Ends of group pairs and non-positively curved cube complexes”, Proc. London Math. Soc. (3) 71 (1995), no. 3, p. 585–617.
  • [Sal87] M. Salvetti – “Topology of the complement of real hyperplanes in 𝐂N{\bf C}^{N}”, Invent. Math. 88 (1987), no. 3, p. 603–618.
  • [Var15] O. Varghese – “A condition that prevents groups from acting fixed point free on cube complexes”, (2015), arXiv:1508.06430.
  • [Wis] D. T. Wise – “The cubical route to understanding groups”, in Proceedings of the International Congress of Mathematicians, (Seoul, 2014), p. 1075–1099.
  • [WW17] D. T. Wise & D. J. Woodhouse – “A cubical flat torus theorem and the bounded packing property”, Israel J. Math. 217 (2017), no. 1, p. 263–281.

Thomas Haettel

Institut Montpelliérain Alexander Grothendieck

CNRS, Univ. Montpellier

thomas.haettel@umontpellier.fr