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

Linearity of Generalized Cactus Groups

Runze Yu Department of Mathematics, University of California, Los Angeles [email protected]
Abstract.

Cactus groups are traditionally defined based on symmetric groups, and pure cactus groups are particular subgroups of cactus groups. Mostovoy [Mos19] showed that pure cactus groups embed into right-angled Coxeter groups. We generalize this result to cactus groups associated with arbitrary finite Coxeter groups and we investigate some representations of generalized cactus groups and deduce the linearity of generalized cactus groups.

1. Introduction

For an integer n>0,n>0, the cactus group JnJ_{n} is the group generated by sp,qs_{p,q} for 1p<qn1\leq p<q\leq n with relations

sp,q2=1,\displaystyle s_{p,q}^{2}=1,
sp,qsm,r=sm,rsp,q,\displaystyle s_{p,q}s_{m,r}=s_{m,r}s_{p,q}, if [p,q][m,r]=\displaystyle\text{if }[p,q]\cap[m,r]=\varnothing
sp,qsm,r=sp+qr,p+qmsp,q,\displaystyle s_{p,q}s_{m,r}=s_{p+q-r,p+q-m}s_{p,q}, if [p,q][m,r]\displaystyle\text{if }[p,q]\supset[m,r]

There is a homomorphism g:JnSng:J_{n}\to S_{n} given by sp,qσp,q,s_{p,q}\mapsto\sigma_{p,q}, where σp,q\sigma_{p,q} reverses the order of elements p,,qp,\cdots,q and leaves the rest unchanged. The kernel of gg is the pure cactus group Γn.\Gamma_{n}.

The subject of cactus groups has generated much interest. Mostovoy showed that the pure cactus group Γn\Gamma_{n} embeds into the diagram group Dn,D_{n}, a right-angled Coxeter group, and Γn\Gamma_{n} is residually nilpotent, see [Mos19]. Davis, Januszkiewicz, and Scott studied the connection between cactus groups and the fundamental groups of blow-ups in [DJS03]. Henriques and Kamnitzer introduced actions of cactus groups on tensor products of crystals and coboundary categories similar to the action of braid groups on braided categories, see [HK06]. Losev studied the case of cactus group defined from the Weyl group of a Lie algebra and the action of the cactus group on the Weyl group, see [Los19], and Bonnafé extended Losev’s construction to any Coxeter group instead of a Weyl group, see [Bon15].

In a Coxeter system (W,S),(W,S), a nonempty subset ISI\subset S is connected if the subgroup WIW_{I} of WW generated by II is finite, and there is no way to write I=JK,I=J\cup K, where JJ and KK are disjoint nonempty sets and WI=WJ×WK.W_{I}=W_{J}\times W_{K}. Set (S)\mathcal{F}(S) to be the family of all connected subsets of S.S.

Definition 1.1.

The generalized cactus group CWC_{W} is generated by {γI:I(S)}\{\gamma_{I}:I\in\mathcal{F}(S)\} subject to relations:

γI2=1,\displaystyle\gamma_{I}^{2}=1,
γIγJ=γJγwJ(I),\displaystyle\gamma_{I}\gamma_{J}=\gamma_{J}\gamma_{w_{J}(I)}, ifIJorWIJ=WI×WJ.\displaystyle\mathrm{if}\,\,I\subset J\,\,\mathrm{or}\,\,W_{I\cup J}=W_{I}\times W_{J}.

Here wJw_{J} is the longest element in WJ.W_{J}. There is a surjective homomorphism gW:CWWg_{W}:C_{W}\to W given by γIwI.\gamma_{I}\mapsto w_{I}. The kernel of gWg_{W} is the generalized pure cactus group. Our main result is that the generalized pure cactus group can be embedded into a right-angled Coxeter group. Given a Coxeter group (W,S)(W,S) and (S)\mathcal{F}(S) the family of connected subsets of S,S, we introduce the right-angled Coxeter group 𝕎\mathbb{W} generated by 𝕊={w(WI):wW,I(S)}\mathbb{S}=\{w(W_{I}):w\in W,I\in\mathcal{F}(S)\} associated with a Coxeter matrix

m(W,W′′)={1,if W=W′′2,if WW′′ or W′′W2,if WW′′={1} and W′′CW(W),otherwise.m(W^{\prime},W^{\prime\prime})=\begin{cases}1,&\text{if }W^{\prime}=W^{\prime\prime}\\ 2,&\text{if }W^{\prime}\subset W^{\prime\prime}\text{ or }W^{\prime\prime}\subset W^{\prime}\\ 2,&\text{if }W^{\prime}\cap W^{\prime\prime}=\{1\}\text{ and }W^{\prime\prime}\subset C_{W}(W^{\prime})\\ \infty,&\text{otherwise}.\end{cases}

In this setting, the group WW acts on (𝕎,𝕊)(\mathbb{W},\mathbb{S}) by wgw,w\mapsto g_{w}, where gw(τW)=τwWw1.g_{w}(\tau_{W^{\prime}})=\tau_{wW^{\prime}w^{-1}}. We prove the generalized cactus group CWC_{W} defined on (W,S)(W,S) embeds into 𝕎Aut(𝕎,𝕊),\mathbb{W}\rtimes\operatorname{Aut}(\mathbb{W},\mathbb{S}), and thereby show that

Theorem 1.2.

The generalized pure cactus group defined by the Coxeter system (W,S)(W,S) embeds into the right-angled Coxeter group (𝕎,𝕊).(\mathbb{W},\mathbb{S}).

Finally, we construct two representations of the generalized cactus group. The generalized cactus group is a special case of the group AA of all lifts of the WW-action on the blow-up Σ#\Sigma_{\#} of a reflection tiling Σ\Sigma to its universal cover [DJS03]. We adapt the representation of AA in [DJS03] to construct a representation of generalized cactus groups.

We construct another representation of generalized cactus groups from the geometrical representation of Coxeter groups in light of Theorem 1.2. Furthermore, this representation restricts to a faithful representation on the generalized cactus group. This shows

Theorem 1.3.

Generalized cactus groups are linear groups.

2. Generalized Cactus Groups

2.1. Definition

The notion of cactus group can be generalized by replacing symmetric groups with general Coxeter groups. Geometrically, it is a special case of the group AA introduced by David, Januszkiewicz and Scott as the group of all lifts of the WW-action on the blow-up Σ#\Sigma_{\#} of a reflection tiling Σ\Sigma to its universal cover in [DJS03].

Let (W,S)(W,S) be a finitely generated Coxeter group and (mst)s,tS(m_{st})_{s,t\in S} be its associated Coxeter matrix. A spherical subgroup of WW is a finite subgroup WIW_{I} generated by some IS.I\subset S. We denote its longest element by wIw_{I} following [Hum90]. A nonempty subset ISI\subset S is connected if WIW_{I} is finite and there is no way to write I=JKI=J\cup K such that JJ and KK are nonempty and WI=WJ×WK.W_{I}=W_{J}\times W_{K}. Let (S)𝒫(S)\mathcal{F}(S)\subset\mathcal{P}(S) denote the family of all connected subsets of S.S.

Definition 2.1.

The generalized cactus group CWC_{W} is generated by {γI:I(S)}\{\gamma_{I}:I\in\mathcal{F}(S)\} subject to relations:

γI2=1\displaystyle\gamma_{I}^{2}=1
γIγJ=γJγwJ(I)\displaystyle\gamma_{I}\gamma_{J}=\gamma_{J}\gamma_{w_{J}(I)} ifIJorWIJ=WI×WJ.\displaystyle\mathrm{if}\,\,I\subset J\,\,\mathrm{or}\,\,W_{I\cup J}=W_{I}\times W_{J}.

Since wJw_{J} is the longest element of WJW_{J}, it follows that wJ(I)w_{J}(I) is still a subset of SS for IJ.I\subset J. In particular if WIJ=WI×WJ,W_{I\cup J}=W_{I}\times W_{J}, then wJ(I)=Iw_{J}(I)=I and γI\gamma_{I} commutes with γJ.\gamma_{J}.

Similar to the case of cactus group, there is a homomorphism gW:CWWg_{W}:C_{W}\to W given by γIwI.\gamma_{I}\mapsto w_{I}. Indeed, wI2=1w_{I}^{2}=1 since it is the longest element in WIW_{I}. If WI×WJ=WIJ,W_{I}\times W_{J}=W_{I\cup J}, then wIw_{I} commutes with wJw_{J}. When IJ,I\subset J, there is wJ(I)Jw_{J}(I)\subset J and the longest element in WwJ(I)=wJ(WI)W_{w_{J}(I)}=w_{J}(W_{I}) is wwJ(I)=wJwIwJ.w_{w_{J}(I)}=w_{J}w_{I}w_{J}. Call the kernel of gWg_{W} the generalized pure cactus group PCW.PC_{W}.

Proposition 2.2.

When WW has type An1,A_{n-1}, there is an isomorphism CWJn.C_{W}\simeq J_{n}.

Proof.

When WW has type An1,A_{n-1}, it is isomorphic to the Coxeter system (Sn,S)(S_{n},S) where S={si:1in1}S=\{s_{i}:1\leq i\leq n-1\} associated by an (n1)×(n1)(n-1)\times(n-1) Coxeter matrix (mij)(m_{ij}) given by

mij={1,i=j,2,|ij|2,3,|ij|=1.m_{ij}=\begin{cases}1,&i=j,\\ 2,&|i-j|\geq 2,\\ 3,&|i-j|=1.\end{cases}

Identify subsets of {1,,n1}\{1,\cdots,n-1\} with subsets of SS via I{si:iI}I\mapsto\{s_{i}:i\in I\} and let Cn:=CWC_{n}:=C_{W} be the generalized cactus group on (W,S).(W,S). If JJ and KK are subsets of {1,2,,n1}\{1,2,\cdots,n-1\} where xy2x-y\geq 2 for all xJx\in J and yK,y\in K, then WJWK=W_{J}\cap W_{K}=\varnothing and every generator of WJW_{J} commutes with every generator in WK.W_{K}. Hence JK(S)J\cup K\notin\mathcal{F}(S) and it follows that (S)\mathcal{F}(S) consists of connected intervals [p,q][p,q] where 1pqn1.1\leq p\leq q\leq n-1.

Let JnJ_{n} be the cactus group generated by sp,qs_{p,q} for 1p<qn.1\leq p<q\leq n. We claim that the map φ:JnCn\varphi:J_{n}\to C_{n} given by sp,qγ[p,q1]s_{p,q}\mapsto\gamma_{[p,q-1]} is a group isomorphism.

We show first that φ\varphi is a homomorphism. Firstly there is γ[p,q1]2=1.\gamma^{2}_{[p,q-1]}=1. If [p,q][m,r]=,[p,q]\cap[m,r]=\varnothing, then by the above discussion,

γ[p,q1]γ[m,r1]=γ[m,r1]γ[p,q1].\gamma_{[p,q-1]}\gamma_{[m,r-1]}=\gamma_{[m,r-1]}\gamma_{[p,q-1]}.

If [p,q][m,r],[p,q]\supset[m,r], then the longest element generated by sp,,sq1s_{p},\cdots,s_{q-1} is σp,q.\sigma_{p,q}. Then

γ[m,r1]γ[p,q1]=γ[p,q1]γσp,q([m,r1])=γ[p,q1]γ[p+qr,p+qm1].\gamma_{[m,r-1]}\gamma_{[p,q-1]}=\gamma_{[p,q-1]}\gamma_{\sigma_{p,q}([m,r-1])}=\gamma_{[p,q-1]}\gamma_{[p+q-r,p+q-m-1]}.

Conjugating by γ[p,q1]\gamma_{[p,q-1]} gives the corresponding relations in JnJ_{n} and therefore φ\varphi is indeed a group homomorphism.

Define ψ:CnJn\psi:C_{n}\to J_{n} by γ[p,q1]sp,q.\gamma_{[p,q-1]}\mapsto s_{p,q}. We have sp,q2=1.s_{p,q}^{2}=1. If W[p,q1]×W[m,r1]=W[p,q1][m,r1],W_{[p,q-1]}\times W_{[m,r-1]}=W_{[p,q-1]\cup[m,r-1]}, then every element in [p,q1][p,q-1] commutes with every element in [m,r1].[m,r-1]. Since they are connected intervals, this implies [p,q][m,r]=[p,q]\cap[m,r]=\varnothing and then sp,qsm,r=sm,rsp,q.s_{p,q}s_{m,r}=s_{m,r}s_{p,q}. If [p,q1][m,r1][p,q-1]\supset[m,r-1], then the relation in CnC_{n} corresponds to the relation sp,qsm,r=sp+qr,p+qmsp,qs_{p,q}s_{m,r}=s_{p+q-r,p+q-m}s_{p,q} in Jn,J_{n}, which is valid. It follows that ψ\psi is also a group homomorphism. Hence CnC_{n} is isomorphic to JnJ_{n} because φψ=idCn\varphi\circ\psi=\mathrm{id}_{C_{n}} and ψφ=idJn.\psi\circ\varphi=\mathrm{id}_{J_{n}}.

2.2. Embedment of the Generalized Pure Cactus Group

Our main theorem in this section is that the generalized pure cactus group can be embedded into a right-angled Coxeter group, generalizing the approach taken by Mostovoy in [[]Proposition 2]Mos:

Theorem 2.3.

Suppose CWC_{W} is the generalized cactus group defined on the Coxeter system (W,S).(W,S). Then the generalized pure cactus group PCWPC_{W} embeds into a right-angled Coxeter group.

We set up a technical lemma before giving the proof. Let (W,S)(W,S) be a right-angled Coxeter group and GG a subgroup of Aut(W,S).\operatorname{Aut}(W,S). Let II be a subset of SS and {gi}iI\{g_{i}\}_{i\in I} a family of involutions in GG satisfying

  1. (1)

    gi(s)Sg_{i}(s)\in S for all iIi\in I and sS.s\in S.

  2. (2)

    Given i,jIi,j\in I with ij=ji,ij=ji, either

    gi(j)I,gj(i)=i, and ggi(j)=gigjgi,g_{i}(j)\in I,g_{j}(i)=i,\text{ and }g_{g_{i}(j)}=g_{i}g_{j}g_{i},

    or

    gj(i)I,gi(j)=j, and ggj(i)=gjgigj.g_{j}(i)\in I,g_{i}(j)=j,\text{ and }g_{g_{j}(i)}=g_{j}g_{i}g_{j}.
  3. (3)

    gi(i)=ig_{i}(i)=i for all iI.i\in I.

Let HH be the subgroup of WGW\rtimes G generated by {igi}iI.\{ig_{i}\}_{i\in I}. Let LL be the group generated by lil_{i} for all iIi\in I subject to relations

  1. (1)

    lilj=lgi(j)lil_{i}l_{j}=l_{g_{i}(j)}l_{i} when ij=ji,gj(i)=i,gi(j)I,ij=ji,g_{j}(i)=i,g_{i}(j)\in I, and ggi(j)=gigjgi,g_{g_{i}(j)}=g_{i}g_{j}g_{i},

  2. (2)

    li2=1l_{i}^{2}=1 for all iI.i\in I.

Lemma 2.4.

The map φ:LH\varphi:L\to H defined by liigil_{i}\mapsto ig_{i} is a group isomorphism.

Proof.

We show first φ\varphi is a group homomorphism. Because gig_{i} are involutions, there is (igi)2=i2gi2=1.(ig_{i})^{2}=i^{2}g_{i}^{2}=1. Now suppose ij=ji,gi(j)I,gj(i)=i,ij=ji,g_{i}(j)\in I,g_{j}(i)=i, and ggi(j)=gigjgi.g_{g_{i}(j)}=g_{i}g_{j}g_{i}. Then

(gi(j)ggi(j))(igi)=(gi(j)(gigjgi)(i))(gigjgigi)=gi(ji)(gigj)=gi(ij)(gigj)=(igi(j))(gigj)=(igi)(jgj).(g_{i}(j)g_{g_{i}(j)})(ig_{i})=(g_{i}(j)(g_{i}g_{j}g_{i})(i))(g_{i}g_{j}g_{i}g_{i})=g_{i}(ji)(g_{i}g_{j})=g_{i}(ij)(g_{i}g_{j})=(ig_{i}(j))(g_{i}g_{j})=(ig_{i})(jg_{j}).

Hence φ\varphi is indeed a group homomorphism. It is also surjective as {igi}iI\{ig_{i}\}_{i\in I} generates H.H. To show injectivity, let i1,i2,,imIi_{1},i_{2},\cdots,i_{m}\in I and let ir=(gi1gi2gir1)(ir)S.i_{r}^{\prime}=(g_{i_{1}}g_{i_{2}}\cdots g_{i_{r-1}})(i_{r})\in S. Note that iri_{r}^{\prime} only depends on iki_{k} for 1kr.1\leq k\leq r. We shall give a lemma first:

Lemma 2.5.

Let li1liml_{i_{1}}\cdots{l_{i_{m}}} be a word in L.L.

  1. (1)

    If there is an rr such that ir=ir+1,i_{r}^{\prime}=i_{r+1}^{\prime}, then there is another word lj1lj2ljm2=li1liml_{j_{1}}l_{j_{2}}\cdots l_{j_{m-2}}=l_{i_{1}}\cdots{l_{i_{m}}} such that jk=ikj_{k}^{\prime}=i_{k}^{\prime} for 1kr11\leq k\leq r-1 and jk=ik+2j_{k}^{\prime}=i_{k+2}^{\prime} for rkm2.r\leq k\leq m-2.

  2. (2)

    If iri_{r}^{\prime} commutes with ir+1,i_{r+1}^{\prime}, then there is another word lj1lj2ljm2=li1liml_{j_{1}}l_{j_{2}}\cdots l_{j_{m-2}}=l_{i_{1}}\cdots{l_{i_{m}}} such that jk=ikj_{k}^{\prime}=i_{k}^{\prime} for kr,r+1,jr=ir+1k\neq r,r+1,j_{r}^{\prime}=i_{r+1}^{\prime} and jr+1=ir.j_{r+1}^{\prime}=i_{r}^{\prime}.

Proof.

Assume ir=ir+1.i_{r}^{\prime}=i_{r+1}^{\prime}. Then ir=gir(ir+1)i_{r}=g_{i_{r}}(i_{r+1}) and then ir=ir+1i_{r}=i_{r+1} since the gg’s are bijective involutions and iri_{r} is fixed by gir+1.g_{i_{r+1}}. Then take jk=ikj_{k}=i_{k} for 1kr11\leq k\leq r-1 and jk=ik+2j_{k}=i_{k+2} for rkm2.r\leq k\leq m-2. Then

lj1lj2ljm2=li1lir1lir+2lim.l_{j_{1}}l_{j_{2}}\cdots l_{j_{m-2}}=l_{i_{1}}\cdots l_{i_{r-1}}l_{i_{r+2}}\cdots l_{i_{m}}.

Moreover, jk=ikj_{k}^{\prime}=i_{k}^{\prime} for 1kr11\leq k\leq r-1 and jk=(gi1gi2gir1gir+2gik1)(ik+2)=ik+2j_{k}^{\prime}=(g_{i_{1}}g_{i_{2}}\cdots g_{i_{r-1}}g_{i_{r+2}}\cdots g_{i_{k-1}})(i_{k+2})=i^{\prime}_{k+2} since gir=gir+1g_{i_{r}}=g_{i_{r+1}} and they are involutions.

Now assume iri_{r}^{\prime} and ir+1i_{r+1}^{\prime} commute and then irgir(ir+1)=gir(ir+1)ir.i_{r}g_{i_{r}}(i_{r+1})=g_{i_{r}}(i_{r+1})i_{r}. Hence iri_{r} commutes with ir+1i_{r+1} after applying gir.g_{i_{r}}. Then either gir(ir+1)Ig_{i_{r}}(i_{r+1})\in I or gir+1(ir)I.g_{i_{r+1}}(i_{r})\in I. If the former is true, then lirlir+1=lgir(ir+1)lirl_{i_{r}}l_{i_{r+1}}=l_{g_{i_{r}}(i_{r+1})}l_{i_{r}} and take

lj1lj2ljm=li1lir1lgir(ir+1)lirlim.l_{j_{1}}l_{j_{2}}\cdots l_{j_{m}}=l_{i_{1}}\cdots l_{i_{r-1}}l_{g_{i_{r}}(i_{r+1})}l_{i_{r}}\cdots{l_{i_{m}}}.

Then there are jk=ikj_{k}^{\prime}=i_{k}^{\prime} for 1kr1,1\leq k\leq r-1, and jk=ikj_{k}^{\prime}=i_{k}^{\prime} for kr+2k\geq r+2 since ggir(ir+1)gir=girgir+1.g_{g_{i_{r}}(i_{r+1})}g_{i_{r}}=g_{i_{r}}g_{i_{r+1}}. Finally, there is jr=ir+1j_{r}^{\prime}=i_{r+1}^{\prime} and jr+1=jrj_{r+1}^{\prime}=j_{r}^{\prime} since ggir(ir+1)gir=girgir+1g_{g_{i_{r}}(i_{r+1})}g_{i_{r}}=g_{i_{r}}g_{i_{r+1}} and iri_{r} is fixed under gir+1.g_{i_{r+1}}. Similarly, if the latter is true, take

lj1lj2ljm=li1lir1lir+1lgir+1(ir)lim.l_{j_{1}}l_{j_{2}}\cdots l_{j_{m}}=l_{i_{1}}\cdots l_{i_{r-1}}l_{i_{r+1}}l_{g_{i_{r+1}}(i_{r})}{l_{i_{m}}}.

This proves Lemma 2.5. ∎

Now suppose φ(li1li2lim)=e.\varphi(l_{i_{1}}l_{i_{2}}\cdots l_{i_{m}})=e. Then i1i2in=e.i_{1}^{\prime}i_{2}^{\prime}\cdots i_{n}^{\prime}=e. Since i1i2ini_{1}^{\prime}i_{2}^{\prime}\cdots i_{n}^{\prime} is a word in the right-angled Coxeter system (W,S),(W,S), there is a sequence of tuples of elements of SS

K1=(K11,K21,,Knd1),,Kd=(K1d,K2d,,Kndd)K^{1}=(K^{1}_{1},K^{1}_{2},\cdots,K^{1}_{n_{d}}),\cdots,K^{d}=(K^{d}_{1},K^{d}_{2},\cdots,K^{d}_{n_{d}})

such that

K1=(K11,K21,,Knd1)=(i1,i2,,im)K^{1}=(K^{1}_{1},K^{1}_{2},\cdots,K^{1}_{n_{d}})=(i_{1}^{\prime},i_{2}^{\prime},\cdots,i_{m}^{\prime})

and KdK^{d} is the empty tuple representing e,e, such that for each s,s, either there exists an rr such that Krs=Kr+1sK^{s}_{r}=K^{s}_{r+1} and

(1) Ks+1=(K1s,,Kr1s,Kr+2s,,Knss)K^{s+1}=(K^{s}_{1},\cdots,K^{s}_{r-1},K^{s}_{r+2},\cdots,K^{s}_{n_{s}})

or there exists a tt such that KtsK^{s}_{t} and Kt+1sK^{s}_{t+1} commute and

(2) Ks+1=(K1s,,Kt+1s,Kts,,Knss).K^{s+1}=(K^{s}_{1},\cdots,K^{s}_{t+1},K^{s}_{t},\cdots,K^{s}_{n_{s}}).

Analyze s=1s=1 as an example. If (1) holds for s=1,s=1, then ir=ir+1i_{r}^{\prime}=i_{r+1}^{\prime} for some rr and by part 1 of Lemma 2.5, K2K^{2} corresponds to some lj1lj2ljm2=li1lim.l_{j_{1}}l_{j_{2}}\cdots l_{j_{m-2}}=l_{i_{1}}\cdots{l_{i_{m}}}. If (2) holds for s=1,s=1, then iri_{r}^{\prime} and ir+1i_{r+1}^{\prime} commute for some r,r, and K2K^{2} corresponds to some lj1lj2ljm=li1liml_{j_{1}}l_{j_{2}}\cdots l_{j_{m}}=l_{i_{1}}\cdots{l_{i_{m}}} by part 2 of Lemma 2.5. By using the same argument inductively, we can conclude that li1lim=e.l_{i_{1}}\cdots{l_{i_{m}}}=e. Hence φ\varphi is an isomorphism. This proves Lemma 2.4. ∎

Now we give the proof of Theorem 2.3:

Proof.

Define 𝕊:={w(WI):I(S) and wW}.\mathbb{S}:=\{w(W_{I}):I\in\mathcal{\mathcal{F}}(S)\text{ and }w\in W\}. Let 𝕄\mathbb{M} be the Coxeter matrix whose entries are given by m(W,W′′){}m(W^{\prime},W^{\prime\prime})\in\mathbb{N}\cup\{\infty\} for each pair (W,W′′)𝕊×𝕊(W^{\prime},W^{\prime\prime})\in\mathbb{S}\times\mathbb{S} where

m(W,W′′)={1,if W=W′′2,if WW′′ or W′′W2,if WW′′={1} and W′′CW(W),otherwisem(W^{\prime},W^{\prime\prime})=\begin{cases}1,&\text{if }W^{\prime}=W^{\prime\prime}\\ 2,&\text{if }W^{\prime}\subset W^{\prime\prime}\text{ or }W^{\prime\prime}\subset W^{\prime}\\ 2,&\text{if }W^{\prime}\cap W^{\prime\prime}=\{1\}\text{ and }W^{\prime\prime}\subset C_{W}(W^{\prime})\\ \infty,&\text{otherwise}\end{cases}

Let (𝕎,𝕊)(\mathbb{W},\mathbb{S}) be the right-angled Coxeter group on generators τW\tau_{W^{\prime}} associated to the Coxeter matrix 𝕄.\mathbb{M}. For each wW,w\in W, there is a group automorphism gwAut(𝕎,𝕊)g_{w}\in\operatorname{Aut}(\mathbb{W},\mathbb{S}) generated by τWτwWw1,\tau_{W^{\prime}}\mapsto\tau_{wW^{\prime}w^{-1}}, since conjugation does not affect inclusion relations between subsets of WW and commutativity relations between elements of W.W. Given I(S),I\in\mathcal{F}(S), put gI=gWI=gwI.g_{I}=g_{W_{I}}=g_{w_{I}}. Note that gIg_{I} is an involution as wIw_{I} is.

Lemma 2.6.

Let I,J(S).I,J\in\mathcal{F}(S). We have

  1. (1)

    gI(τWI)=τWI.g_{I}(\tau_{W_{I}})=\tau_{W_{I}}.

  2. (2)

    If IJ,I\subset J, then gI(τWJ)=τWJ,gJ(τWI)=τWwJ(I),g_{I}(\tau_{W_{J}})=\tau_{W_{J}},g_{J}(\tau_{W_{I}})=\tau_{W_{w_{J}(I)}}, and gwJ(I)=gJgIgJ.g_{w_{J}(I)}=g_{J}g_{I}g_{J}.

  3. (3)

    If WIJ=WI×WJ,W_{I\cup J}=W_{I}\times W_{J}, then gI(τWJ)=τWJ,gJ(τWI)=τWI,gwJ(I)=gJgIgJ,g_{I}(\tau_{W_{J}})=\tau_{W_{J}},g_{J}(\tau_{W_{I}})=\tau_{W_{I}},g_{w_{J}(I)}=g_{J}g_{I}g_{J}, and gwI(J)=gIgJgI.g_{w_{I}(J)}=g_{I}g_{J}g_{I}.

  4. (4)

    m(WI,WJ)=2,gJ(τWI)=τWIm(W_{I},W_{J})=2,g_{J}(\tau_{W_{I}})=\tau_{W_{I}} and gwI(J)=gIgJgIg_{w_{I}(J)}=g_{I}g_{J}g_{I} if and only if JIJ\subset I or WI×WJ=WIJW_{I}\times W_{J}=W_{I\cup J} or IJI\subset J with wJ(I)=I.w_{J}(I)=I.

Proof.

(1) is clear. For (2), assume IJ.I\subset J. Then wIWIWJw_{I}\in W_{I}\subset W_{J} and then wI(WJ)=WJw_{I}(W_{J})=W_{J} since conjugation is bijective. Hence gI(τWJ)=τWJ.g_{I}(\tau_{W_{J}})=\tau_{W_{J}}. Furthermore, because WIW_{I} is generated by I,I, a generating set of wJ(WI)w_{J}(W_{I}) is wJ(I)wJ(J)=Jw_{J}(I)\subset w_{J}(J)=J since wJw_{J} is the longest element. Then gJ(τWI)=τWwJ(I).g_{J}(\tau_{W_{I}})=\tau_{W_{w_{J}(I)}}. Finally, gwJ(I)g_{w_{J}(I)} is conjugation by the longest element in wJ(WI),w_{J}(W_{I}), which is wJ(wI)=wJwIwJ.w_{J}(w_{I})=w_{J}w_{I}w_{J}. It follows immediately that gwJ(I)=gJgIgJ.g_{w_{J}(I)}=g_{J}g_{I}g_{J}.

(3) can be shown by two similar arguments in which wJ(WI)=WIw_{J}(W_{I})=W_{I} and wI(WJ)=WJw_{I}(W_{J})=W_{J} respectively.

Sufficiency in (4) follows from (2) and (3). Now assume m(WI,WJ)=2,gJ(τWI)=τWIm(W_{I},W_{J})=2,g_{J}(\tau_{W_{I}})=\tau_{W_{I}} and gwI(J)=gIgJgI.g_{w_{I}(J)}=g_{I}g_{J}g_{I}. The only nontrivial case is IJI\subset J by definition of the Coxeter matrix M.M. In this case, (2) implies WI=WwJ(I)=wJ(WI).W_{I}=W_{w_{J}(I)}=w_{J}(W_{I}). In particular, wJ(I)=wJ(WI)S=WIS=I.w_{J}(I)=w_{J}(W_{I})\cap S=W_{I}\cap S=I. This completes the proof. ∎

We conclude the proof of theorem 2.3 by showing the following lemma:

Lemma 2.7.

There is a monomorphism CW𝕎Aut(𝕎,𝕊)C_{W}\hookrightarrow\mathbb{W}\rtimes\operatorname{Aut}(\mathbb{W},\mathbb{S}) defined by γIτWIgI.\gamma_{I}\mapsto\tau_{W_{I}}\cdot g_{I}. It restricts to a monomorphism PCW𝕎.PC_{W}\hookrightarrow\mathbb{W}.

Proof.

There is a monomorphism h:L𝕎Aut(𝕎,𝕊),h:L\to\mathbb{W}\rtimes\operatorname{Aut}(\mathbb{W},\mathbb{S}), where LCWL\simeq C_{W} is the group generated by lIl_{I} as in Lemma 2.4. This proves the first part.

To show the second part, observe that the following diagram commutes:

CW{C_{W}}𝕎Aut(𝕎,𝕊){\mathbb{W}\rtimes\operatorname{Aut}(\mathbb{W},\mathbb{S})}W{W}Aut(𝕎,𝕊){\operatorname{Aut}(\mathbb{W},\mathbb{S})}gWg_{W}hhggProj

in which

Proj(h(γI))=Proj(τWIgI)=gI,\mathrm{Proj}(h(\gamma_{I}))=\mathrm{Proj}(\tau_{W_{I}}\cdot g_{I})=g_{I},
g(gW(γI))=g(wI)=gwI=gI.g(g_{W}(\gamma_{I}))=g(w_{I})=g_{w_{I}}=g_{I}.

Because hh is injective, the kernel of gWg_{W} lies inside the kernel of the projection, which is isomorphic to 𝕎.\mathbb{W}. This completes the proof. ∎

Theorem 2.3 now follows immediately from Lemma 2.7. ∎

As an example, the usual cactus group JnJ_{n} defined on the Coxeter group SnS_{n} coincides with the corresponding generalized cactus group Cn.C_{n}. It follows from Theorem 2.3 that JnJ_{n} embeds into DnSnD_{n}\rtimes S_{n} where DnD_{n} is the diagram group and CnC_{n} embeds into 𝕎Aut(𝕎,𝕊)\mathbb{W}\rtimes\operatorname{Aut}(\mathbb{W},\mathbb{S}) for an appropriate Coxeter system (𝕎,𝕊),(\mathbb{W},\mathbb{S}), see [Mos19]. Furthermore, there is a bijection ff from 𝕊\mathbb{S} to the set of subsets I{1,2,,n}I\subset\{1,2,\cdots,n\} with |I|2|I|\geq 2 given by

W{si,si+1,,sj}{i,i+1,,j+1}W_{\{s_{i},s_{i+1},\cdots,s_{j}\}}\mapsto\{i,i+1,\cdots,j+1\}

and

wWIw1w(f(WI)).wW_{I}w^{-1}\mapsto w(f(W_{I})).

It follows that there is a group isomorphism 𝕎Dn\mathbb{W}\simeq D_{n} and a commutative diagram

CW{C_{W}}𝕎Aut(𝕎,𝕊){\mathbb{W}\rtimes\operatorname{Aut}(\mathbb{W},\mathbb{S})}Jn{J_{n}}DnSn{D_{n}\rtimes S_{n}}\sim

3. Representations of Generalized Cactus Groups

3.1. Geometrical Representation of Coxeter Groups

Let (W,S)(W,S) be a Coxeter system and MM its associated Coxeter matrix. A representation of WAut(W,S)W\rtimes\operatorname{Aut}(W,S) can be constructed as follows. Let E=SE=\mathbb{R}^{S} with standard basis {εs}sS.\{\varepsilon_{s}\}_{s\in S}. Define a bilinear form BtB_{t} on EE by

Bt(εs,εv)={cos(π/ms,v)ms,v<tms,v=B_{t}(\varepsilon_{s},\varepsilon_{v})=\begin{cases}-\cos\left(\pi/m_{s,v}\right)&m_{s,v}<\infty\\ -t&m_{s,v}=\infty\end{cases}

For any t>0,t>0, there is a representation π\pi of WW (in particular the Tits representation when t=1t=1) generated by sσss\mapsto\sigma_{s} where

σs(x)=x2Bt(x,εs)εs.\sigma_{s}(x)=x-2B_{t}(x,\varepsilon_{s})\varepsilon_{s}.

To show π\pi is a representation, it suffices to show σsσv\sigma_{s}\sigma_{v} has order ms,vm_{s,v} for s,vS.s,v\in S. Clearly σs\sigma_{s} preserves the bilinear form Bt.B_{t}. Let λ=Bt(es,ev).\lambda=B_{t}(e_{s},e_{v}). Then

(σsσv)(es)=σs(es2λev)=es2λ(ev2λes)=(4λ21)es2λev(\sigma_{s}\sigma_{v})(e_{s})=\sigma_{s}(e_{s}-2\lambda e_{v})=-e_{s}-2\lambda(e_{v}-2\lambda e_{s})=(4\lambda^{2}-1)e_{s}-2\lambda e_{v}

and (σsσv)(ev)=ev+2λes.(\sigma_{s}\sigma_{v})(e_{v})=-e_{v}+2\lambda e_{s}. If ms,v=,m_{s,v}=\infty, then λ=t.\lambda=-t. For all t1,t\geq 1, the equation k2+2λk+1=0k^{2}+2\lambda k+1=0 has a positive real root k=k0>0.k=k_{0}>0. Set r=4λ2+2k0λ1r=4\lambda^{2}+2k_{0}\lambda-1 and assume r0r\neq 0 (this is true for all but finitely many tt). Then

(σsσv)(es+k0ev)=(4λ2+2λk01)es(2λ+k0)ev=r(es+k0ev).(\sigma_{s}\sigma_{v})(e_{s}+k_{0}e_{v})=(4\lambda^{2}+2\lambda k_{0}-1)e_{s}-(2\lambda+k_{0})e_{v}=r(e_{s}+k_{0}e_{v}).

Hence using the second equation which can be rewritten into (σsσv)(ev)=2λ(es+k0ev)(2λk0+1)ev,(\sigma_{s}\sigma_{v})(e_{v})=2\lambda(e_{s}+k_{0}e_{v})-(2\lambda k_{0}+1)e_{v}, we can obtain

(σsσv)m(ev)=am(es+k0ev)+(2λk01)mev(\sigma_{s}\sigma_{v})^{m}(e_{v})=a_{m}(e_{s}+k_{0}e_{v})+(-2\lambda k_{0}-1)^{m}e_{v}

where a1=2λa_{1}=2\lambda and am+1=ram+2λ(2λk01)m.a_{m+1}=ra_{m}+2\lambda(-2\lambda k_{0}-1)^{m}. If t>1,t>1, then 2λk01=k02>1-2\lambda k_{0}-1=k_{0}^{2}>1 and (σsσv)m(\sigma_{s}\sigma_{v})^{m} can never be the identity. If t=1,t=1, then am=2ma_{m}=2m and (σsσv)m(\sigma_{s}\sigma_{v})^{m} cannot be the identity, either. Thus σsσv\sigma_{s}\sigma_{v} has order .\infty.

When ms,v<,m_{s,v}<\infty, we use the identification span(es,ev)\text{span}(e_{s},e_{v})\approx\mathbb{C} by identifying ese_{s} to 1 and eve_{v} to eiθ,-e^{-i\theta}, where θ=π/ms,v.\theta=\pi/m_{s,v}. Then σsσv\sigma_{s}\sigma_{v} acts by a rotation of 2π/ms,v2\pi/m_{s,v} and thus has order ms,v.m_{s,v}. Hence π\pi is a representation.

Moreover, a faithful representation π\pi^{\prime} of Aut(W,S)\operatorname{Aut}(W,S) on GL(E)GL(E) is given by π(g)(εs)=εg(s)\pi^{\prime}(g)(\varepsilon_{s})=\varepsilon_{g(s)} for all basis vectors εs.\varepsilon_{s}. It is possible to construct a representation of WAut(W,S)W\rtimes\operatorname{Aut}(W,S) from π\pi and π.\pi^{\prime}.

Proposition 3.1.

Let Π:WAut(W,S)GL(E)\Pi:W\rtimes\operatorname{Aut}(W,S)\to GL(E) be given by Π(τg)=π(τ)π(g).\Pi(\tau g)=\pi(\tau)\pi^{\prime}(g). Then Π\Pi is a representation of WAut(W,S).W\rtimes\operatorname{Aut}(W,S).

Proof.

Since gg is an automorphism and preserves the Coxeter matrix, π(g)\pi^{\prime}(g) preserves the bilinear form Bt.B_{t}. As π\pi and π\pi^{\prime} are representations of WW and Aut(W,S)\operatorname{Aut}(W,S) respectively, we need to check Π\Pi respects the product structure induced by the semi-direct product. It suffices to check that for any gAut(W,S)g\in\operatorname{Aut}(W,S) and vS,v\in S,

π(g)π(v)=π(g(v))π(g).\pi^{\prime}(g)\pi(v)=\pi(g(v))\pi^{\prime}(g).

In fact, for any basis vector εu,\varepsilon_{u}, there is

(π(g)π(v))(εu)\displaystyle(\pi^{\prime}(g)\circ\pi(v))(\varepsilon_{u}) =π(g)(εu2Bt(εu,εv)εv)\displaystyle=\pi^{\prime}(g)(\varepsilon_{u}-2B_{t}(\varepsilon_{u},\varepsilon_{v})\varepsilon_{v})
=εg(u)2Bt(εg(u),εg(v))εg(v)\displaystyle=\varepsilon_{g(u)}-2B_{t}(\varepsilon_{g(u)},\varepsilon_{g(v)})\varepsilon_{g(v)}
=π(τg(v))(εg(u))\displaystyle=\pi(\tau_{g(v)})(\varepsilon_{g(u)})
=(π(τg(v))π(g))(εu).\displaystyle=(\pi(\tau_{g(v)})\circ\pi^{\prime}(g))(\varepsilon_{u}).

Hence Π\Pi is a representation. ∎

When t=1,t=1, the bilinear form is the canonical bilinear form BB induced by the Coxeter matrix M.M. The following result is classical but we could not find a reference. The injectivity of π\pi is for example in [[]Corollary 5.4]Hum.

Lemma 3.2.

When t=1,t=1, the representation Π\Pi is faithful.

Proof.

Suppose Π(wg)=id.\Pi(wg)=\mathrm{id}. If we,w\neq e, then there exists sSs\in S such that l(ws)<l(w).l(ws)<l(w). Let s=g1(s).s^{\prime}=g^{-1}(s). Then [[]Theorem 5.4]Hum shows that B(w(εs),εt)0B(w(\varepsilon_{s}),\varepsilon_{t})\leq 0 for all tS.t\in S. Then B(wg(εs),εt)0.B(wg(\varepsilon_{s^{\prime}}),\varepsilon_{t})\leq 0. But B(wg(εs),εs)=B(εs,εs)=1,B(wg(\varepsilon_{s^{\prime}}),\varepsilon_{s^{\prime}})=B(\varepsilon_{s^{\prime}},\varepsilon_{s^{\prime}})=1, a contradiction. Hence w=ew=e and then g=idg=\mathrm{id} since π\pi^{\prime} is faithful. This shows faithfulness of Π.\Pi.

Now pick a tt so that BtB_{t} is nondegenerate. Theorem 1 shows that CWC_{W} can be embedded into the semi-direct product 𝕎Aut(𝕎,𝕊).\mathbb{W}\rtimes\operatorname{Aut}(\mathbb{W},\mathbb{S}). Define 𝔼\mathbb{E} to be the vector space on basis {εW}W𝕊\{\varepsilon_{W^{\prime}}\}_{W^{\prime}\in\mathbb{S}} over .\mathbb{R}. The space EE embeds into 𝔼\mathbb{E} through the map εIεWI.\varepsilon_{I}\mapsto\varepsilon_{W_{I}}. Through this map, the bilinear form BtB_{t} on EE agrees with the bilinear form B¯t\bar{B}_{t} on 𝔼\mathbb{E} generated by

Bt¯(εW,εW′′)={cos(π/m(W,W′′)),m(W,W′′)<t.m(W,W′′)=\bar{B_{t}}(\varepsilon_{W^{\prime}},\varepsilon_{W^{\prime\prime}})=\begin{cases}-\cos\left(\pi/m(W^{\prime},W^{\prime\prime})\right),&m(W^{\prime},W^{\prime\prime})<\infty\\ -t.&m(W^{\prime},W^{\prime\prime})=\infty\end{cases}

Let Π\Pi be the representation of 𝕎Aut(𝕎,𝕊)\mathbb{W}\rtimes\operatorname{Aut}(\mathbb{W},\mathbb{S}) on 𝔼\mathbb{E} constructed previously. By restricting to CWC_{W} and its isomorphic image in 𝕎Aut(𝕎,𝕊),\mathbb{W}\rtimes\operatorname{Aut}(\mathbb{W},\mathbb{S}), Π\Pi becomes a representation of CW.C_{W}.

Corollary 3.3.

The representation of CWC_{W} on 𝔼\mathbb{E} is faithful. Therefore, generalized cactus groups are linear groups.

3.2. Another Linear Representation

This section is motivated by sections 5.10, 6.1, and 6.2 in [DJS03]. The generalized cactus group and certain representations can be constructed from a geometrical viewpoint as in [DJS03]. Here we give a direct algebraic construction. Let (W,S)(W,S) be a Coxeter system and let CWC_{W} be the generalized cactus group on (W,S).(W,S). For I,J(S),I,J\in\mathcal{F}(S), set

m(I,J)={1,if I=J2,if IJ or JI2,if WI×WJ=WIJ,otherwise.m(I,J)=\begin{cases}1,&\text{if }I=J\\ 2,&\text{if }I\subset J\text{ or }J\subset I\\ 2,&\text{if }W_{I}\times W_{J}=W_{I\cup J}\\ \infty,&\text{otherwise.}\end{cases}

Let MM be the Coxeter matrix indexed by (S)×(S)\mathcal{F}(S)\times\mathcal{F}(S) with entries given by m(I,J).m(I,J). A representation of CWC_{W} can be constructed as in [DJS03] . Let EE be the real vector space spanned by a basis {εI}I(S).\{\varepsilon_{I}\}_{I\in\mathcal{F}(S)}. Define

𝒞(I):={J(S):JI}.\mathcal{C}(I):=\{J\in\mathcal{F}(S):J\subset I\}.

Note that I𝒞(I).I\notin\mathcal{C}(I). Let EIEE_{I}\subset E denote the subspace spanned by the set

{εJεJ:J𝒞(I),J=wI(J)}.\{\varepsilon_{J}-\varepsilon_{J^{\prime}}:J\in\mathcal{C}(I),J^{\prime}=w_{I}(J)\}.

For each t,t\in\mathbb{R}, define a symmetric bilinear form on EE by

Bt(εI,εJ)={1,I=J0,m(I,J)=2t,otherwise.B_{t}(\varepsilon_{I},\varepsilon_{J})=\begin{cases}1,&I=J\\ 0,&m(I,J)=2\\ -t,&\text{otherwise.}\end{cases}

B0B_{0} is the standard inner product on E,E, and B1B_{1} is the canonical inner product associated to the Coxeter matrix M.M. It follows that detB0=10\det B_{0}=1\neq 0 and detBt\det B_{t} is a nonzero polynomial in t.t. Hence BtB_{t} is nondegenerate on EE for sufficiently large t.t. Let FIF_{I} be the orthogonal complement of εIEI.\mathbb{R}\varepsilon_{I}\oplus E_{I}. Then we have an orthogonal decomposition

E=εIEIFI.E=\mathbb{R}\varepsilon_{I}\oplus E_{I}\oplus F_{I}.

Now define ρI:EE\rho_{I}:E\to E by

ρI=id|εIid|EIid|FI.\rho_{I}=-\left.\mathrm{id}\right|_{\mathbb{R}\varepsilon_{I}}\oplus-\left.\mathrm{id}\right|_{E_{I}}\oplus\left.\mathrm{id}\right|_{F_{I}}.
Proposition 3.4.

The map γIρI\gamma_{I}\to\rho_{I} extends to a representation of CW.C_{W}.

Proof.

It suffices to show the relations in CWC_{W} still hold with γI\gamma_{I} replaced by ρI:\rho_{I}:

  1. (a)

    ρI2=1,\rho_{I}^{2}=1,

  2. (b)

    ρIρJ=ρJρI\rho_{I}\rho_{J}=\rho_{J}\rho_{I} if WIJ=WI×WJ,W_{I\cup J}=W_{I}\times W_{J}, and

  3. (c)

    ρIρJ=ρJρwJ(I)\rho_{I}\rho_{J}=\rho_{J}\rho_{w_{J}(I)} if IJ.I\subset J.

ρI2=1\rho_{I}^{2}=1 since ρI\rho_{I} is an involution. To show (b) and (c), we give the following lemma:

Lemma 3.5.
  1. (1)

    If WIJ=WI×WJ,W_{I\cup J}=W_{I}\times W_{J}, then ρI(εJ)=εJ\rho_{I}(\varepsilon_{J})=\varepsilon_{J} and ρI|EJ=id.\left.\rho_{I}\right|_{E_{J}}=\mathrm{id}.

  2. (2)

    If J𝒞(I),J\in\mathcal{C}(I), then ρI(εJ)=εJ\rho_{I}(\varepsilon_{J})=\varepsilon_{J^{\prime}} and ρI(EJ)=EJ\rho_{I}(E_{J})=E_{J^{\prime}} where J=wI(J).J^{\prime}=w_{I}(J).

Proof.

For (1), note first that εJ\varepsilon_{J} is orthogonal to εI.\varepsilon_{I}. Moreover, for any L𝒞(I),L\in\mathcal{C}(I), we have WL×WJ=WLJW_{L}\times W_{J}=W_{L\cup J} and the same is true with LL replaced by wI(L).w_{I}(L). Hence εJ\varepsilon_{J} is orthogonal to EI.E_{I}. Then εJFI\varepsilon_{J}\in F_{I} and εJ\varepsilon_{J} is fixed by ρI.\rho_{I}.

Furthermore, suppose K𝒞(J).K\in\mathcal{C}(J). Then WK×WI=WKIW_{K}\times W_{I}=W_{K\cup I} and ρI(εK)=εK\rho_{I}(\varepsilon_{K})=\varepsilon_{K} by the previous argument. Moreover, K=wJ(K)K^{\prime}=w_{J}(K) is also a subset of J,J, and then ρI(εK)=εK\rho_{I}(\varepsilon_{K^{\prime}})=\varepsilon_{K^{\prime}} by the same argument with KK replaced by K.K^{\prime}. This implies both εK\varepsilon_{K} and εK\varepsilon_{K^{\prime}} are fixed by ρI\rho_{I} and then ρI\rho_{I} is identity on EJ.E_{J}.

For (2), if J𝒞(I),J\in\mathcal{C}(I), then J=wI(J)𝒞(I)J^{\prime}=w_{I}(J)\in\mathcal{C}(I) and the vector εJ+εJ\varepsilon_{J}+\varepsilon_{J^{\prime}} is orthogonal to εI.\varepsilon_{I}. Let L𝒞(I)L\in\mathcal{C}(I) and L=wI(L).L^{\prime}=w_{I}(L). We have εJ,εL=εJ,εL\langle\varepsilon_{J},\varepsilon_{L}\rangle=\langle\varepsilon_{J^{\prime}},\varepsilon_{L^{\prime}}\rangle and εJ,εL=εJ,εL.\langle\varepsilon_{J},\varepsilon_{L^{\prime}}\rangle=\langle\varepsilon_{J^{\prime}},\varepsilon_{L}\rangle. It follows that

εJ+εJ,εLεL=0.\langle\varepsilon_{J}+\varepsilon_{J^{\prime}},\varepsilon_{L}-\varepsilon_{L^{\prime}}\rangle=0.

Hence εJ+εJ\varepsilon_{J}+\varepsilon_{J^{\prime}} is orthogonal to EI.E_{I}. Then εJ+εJFI\varepsilon_{J}+\varepsilon_{J^{\prime}}\in F_{I} and

ρI(εJ)=12(ρI(εJ+εJ)+ρI(εJεJ))=12(εJ+εJ(εJεJ))=εJ.\rho_{I}(\varepsilon_{J})=\dfrac{1}{2}\left(\rho_{I}(\varepsilon_{J}+\varepsilon_{J^{\prime}})+\rho_{I}(\varepsilon_{J}-\varepsilon_{J^{\prime}})\right)=\dfrac{1}{2}(\varepsilon_{J}+\varepsilon_{J^{\prime}}-(\varepsilon_{J}-\varepsilon_{J^{\prime}}))=\varepsilon_{J^{\prime}}.

For the second part, suppose K𝒞(J).K\in\mathcal{C}(J). Then K=wJ(K)JIK^{\prime}=w_{J}(K)\subset J\subset I and

ρI(εKεK)=ρI(εK)ρI(εK)=εwI(K)εwI(wJ(K)).\rho_{I}(\varepsilon_{K}-\varepsilon_{K^{\prime}})=\rho_{I}(\varepsilon_{K})-\rho_{I}(\varepsilon_{K^{\prime}})=\varepsilon_{w_{I}(K)}-\varepsilon_{w_{I}(w_{J}(K))}.

Note that wIwJ=wIwJwI1wI=wJwI.w_{I}w_{J}=w_{I}w_{J}w_{I}^{-1}w_{I}=w_{J^{\prime}}w_{I}. Then

ρI(εKεK)=εwI(K)εwI(wJ(K))=εwI(K)εwJ(wI(K))EJ\rho_{I}(\varepsilon_{K}-\varepsilon_{K^{\prime}})=\varepsilon_{w_{I}(K)}-\varepsilon_{w_{I}(w_{J}(K))}=\varepsilon_{w_{I}(K)}-\varepsilon_{w_{J^{\prime}}(w_{I}(K))}\in E_{J^{\prime}}

since wI(K)wI(J)=J.w_{I}(K)\subset w_{I}(J)=J^{\prime}. Hence ρI(EJ)EJ.\rho_{I}(E_{J})\subset E_{J^{\prime}}. A similar argument shows that ρI(EJ)EJ\rho_{I}(E_{J^{\prime}})\subset E_{J} and then ρI(EJ)=EJ\rho_{I}(E_{J})=E_{J^{\prime}} as ρI\rho_{I} is an involution. ∎

Case (b) now follows from the commutative diagram implied by part (1) of the lemma:

(εJEJ)FJ{(\mathbb{R}\varepsilon_{J}\oplus E_{J})\oplus F_{J}}(εJEJ)FJ{(\mathbb{R}\varepsilon_{J}\oplus E_{J})\oplus F_{J}}(εJEJ)FJ{(\mathbb{R}\varepsilon_{J}\oplus E_{J})\oplus F_{J}}(εJEJ)FJ{(\mathbb{R}\varepsilon_{J}\oplus E_{J})\oplus F_{J}}ρJ=idid\rho_{J}=-\mathrm{id}\oplus\mathrm{id}ρI\rho_{I}ρI\rho_{I}ρJ=idid\rho_{J}=-\mathrm{id}\oplus\mathrm{id}

Part (2) of the lemma gives a similar commutative diagram for case (c):

(εJEJ)FJ{(\mathbb{R}\varepsilon_{J}\oplus E_{J})\oplus F_{J}}(εJEJ)FJ{(\mathbb{R}\varepsilon_{J^{\prime}}\oplus E_{J^{\prime}})\oplus F_{J^{\prime}}}(εJEJ)FJ{(\mathbb{R}\varepsilon_{J}\oplus E_{J})\oplus F_{J}}(εJEJ)FJ{(\mathbb{R}\varepsilon_{J^{\prime}}\oplus E_{J^{\prime}})\oplus F_{J^{\prime}}}ρJ=idid\rho_{J}=-\mathrm{id}\oplus\mathrm{id}ρI\rho_{I}ρI\rho_{I}ρJ=idid\rho_{J^{\prime}}=-\mathrm{id}\oplus\mathrm{id}

This completes the proof. ∎

When t=0,t=0, εJ\varepsilon_{J} is orthogonal to εI\varepsilon_{I} and EIE_{I} and ρ(εJ)=εJ.\rho(\varepsilon_{J})=\varepsilon_{J}. Lemma 5 implies the images of the standard basis {εJ}J(S)\{\varepsilon_{J}\}_{J\in\mathcal{F}(S)} under any ρ(w)\rho(w) for wCWw\in C_{W} lie in the set {±εJ}J(S)\{\pm\varepsilon_{J}\}_{J\in\mathcal{F}(S)} when t=0.t=0. Hence the representation ρ:CWGL(E)\rho:C_{W}\to GL(E) cannot be faithful when CWC_{W} is infinite, nor will ρ\rho be faithful when restricted to an infinite subgroup of CWC_{W} (for example the generalized pure cactus group PCWPC_{W}).

3.3. Examples of Generalized Cactus Groups

In this section, we analyze a few examples of generalized cactus groups and their representations.

3.3.1. Type A2A_{2}

The cactus group J3J_{3} is defined on the Coxeter group S3S_{3} and is generated by s1s_{1} and s2s_{2} with

={{s1},{s2},{s1,s2}}.\mathcal{F}=\{\{s_{1}\},\{s_{2}\},\{s_{1},s_{2}\}\}.

Then J3J_{3} can be generated by B=γ{s2}B=\gamma_{\{s_{2}\}} and C=γ{s1,s2}C=\gamma_{\{s_{1},s_{2}\}} and the representation ρ:J3GL(,3)\rho:J_{3}\to GL(\mathbb{R},3) is given by involutions

ρ(B)=(1002t10001),ρ(C)=(010100001)\rho(B)=\begin{pmatrix}1&0&0\\ 2t&-1&0\\ 0&0&1\end{pmatrix},\rho(C)=\begin{pmatrix}0&1&0\\ 1&0&0\\ 0&0&-1\end{pmatrix}

and then ρ|U\left.\rho\right|_{U} is irreducible where U=span(ε1,ε2).U=\mathrm{span}(\varepsilon_{1},\varepsilon_{2}).

Furthermore, the group J3J_{3} embeds into D3AutD3=D3S3D_{3}\rtimes\operatorname{Aut}D_{3}=D_{3}\rtimes S_{3} by Theorem 2.3. Then the representation Π:J3D3S3GL(,4)\Pi:J_{3}\hookrightarrow D_{3}\rtimes S_{3}\to GL(\mathbb{R},4) is given by

Π(B)=(010010002t2t100001),Π(C)=(0010010010000001).\Pi(B)=\begin{pmatrix}0&1&0&0\\ 1&0&0&0\\ 2t&2t&-1&0\\ 0&0&0&1\end{pmatrix},\,\Pi(C)=\begin{pmatrix}0&0&1&0\\ 0&1&0&0\\ 1&0&0&0\\ 0&0&0&-1\end{pmatrix}.

The subspace U1=span(ε1,ε2,ε3)U_{1}=\mathrm{span}(\varepsilon_{1},\varepsilon_{2},\varepsilon_{3}) is invariant under Π\Pi and a stable vector of Π\Pi is v0=ε1ε2+ε3.v_{0}=\varepsilon_{1}-\varepsilon_{2}+\varepsilon_{3}. Hence there is a quotient representation of Π\Pi on the quotient space U1/v0U_{1}/\langle v_{0}\rangle with basis ε¯1\bar{\varepsilon}_{1} and ε¯3.\bar{\varepsilon}_{3}. Under this basis, the quotient representation is given by

B(102t+11),C(0110).B\mapsto\begin{pmatrix}1&0\\ 2t+1&-1\end{pmatrix},\,C\mapsto\begin{pmatrix}0&1\\ 1&0\end{pmatrix}.

The quotient representation is isomorphic to the representation ρ\rho with a change in parameter tt to t+1/2.t+1/2.

3.3.2. Type I2(n)I_{2}(n)

The dihedral group of order nn is generated by {a,b}\{a,b\} with an associated Coxeter matrix

(1nn1)\begin{pmatrix}1&n\\ n&1\end{pmatrix}

and ={{a},{b},{a,b}}.\mathcal{F}=\{\{a\},\{b\},\{a,b\}\}. The longest element in D2nD_{2n} is (ab)n/2(ab)^{n/2} if nn is even and b(ab)(n1)/2b(ab)^{(n-1)/2} if nn is odd.

Let CD2nCD_{2n} be the generalized cactus group on D2nD_{2n} generated by γ1=γa,γ2=γb\gamma_{1}=\gamma_{a},\gamma_{2}=\gamma_{b} and γ3=γab.\gamma_{3}=\gamma_{ab}. Then CD2nCD_{2n} is isomorphic to (2)3(\mathbb{Z}_{2})^{3} when n=2,n=2, to D×2D_{\infty}\times\mathbb{Z}_{2} when n4n\geq 4 and nn is even, and to J3J_{3} (hence to DD_{\infty}) when nn is odd.

Let nn be an odd integer, and then CD2nCD_{2n} admits the same representation ρt\rho_{t} on EE as J3J_{3} does. However, the basis of the vector space 𝔼\mathbb{E} consists of the vectors εi\varepsilon_{i} for 0in,0\leq i\leq n, indexed by the subgroups Ai={e,(ab)ia}A_{i}=\{e,(ab)^{i}a\} for i=0,,n1i=0,\cdots,n-1 (as b=(ab)n1ab=(ab)^{n-1}a is a conjugation of aa) and D2nD_{2n} for i=ni=n. Using the notation in the previous section where B=γbB=\gamma_{b} and C=γab,C=\gamma_{ab}, we have

Π(B)(εi)={εni2+2tεn1,0in2εn1,i=n1εn,i=n\Pi(B)(\varepsilon_{i})=\begin{cases}\varepsilon_{n-i-2}+2t\varepsilon_{n-1},&0\leq i\leq n-2\\ -\varepsilon_{n-1},&i=n-1\\ \varepsilon_{n},&i=n\end{cases}

and

Π(C)(εi)={εn1i,0in1εn,i=n.\Pi(C)(\varepsilon_{i})=\begin{cases}\varepsilon_{n-1-i},&0\leq i\leq n-1\\ -\varepsilon_{n},&i=n.\end{cases}

There are two subrepresentations of Π\Pi spanned by εi\varepsilon_{i} for 0in10\leq i\leq n-1 and εn,\varepsilon_{n}, respectively. A stable subspace VV of Π\Pi is spanned by εiεn2i\varepsilon_{i}-\varepsilon_{n-2-i} for 0in2,0\leq i\leq n-2, and a stable vector of Π\Pi is

v=ε0ε1++ε2k.v=\varepsilon_{0}-\varepsilon_{1}+\cdots+\varepsilon_{2k}.

Acknowledgements

The author would like to warmly thank Professor Raphaël Rouquier for supporting this work. This project was partly funded by NSF Grant DMS-1702305 as part of the 2021 Summer REU Program at UCLA.

References

  • [BB05] Anders Bjorner and Francesco Brenti “Combinatorics of Coxeter Groups” Springer, Berlin, Heidelberg, 2005
  • [BCP97] Wieb Bosma, John Cannon and Catherine Playoust “The Magma algebra system. I. The user language” Computational algebra and number theory (London, 1993) In J. Symbolic Comput. 24.3-4, 1997, pp. 235–265 DOI: 10.1006/jsco.1996.0125
  • [Bon15] Cédric Bonnafé “Cells and Cacti” In International Mathematics Research Notices 2016.19, 2015, pp. 5775–5800 DOI: 10.1093/imrn/rnv324
  • [Dav12] Michael Davis “The Geometry and Topology of Coxeter Groups. (LMS-32):” Princeton University Press, 2012
  • [DJS03] M Davis, T Januszkiewicz and R Scott “Fundamental groups of blow-ups” In Advances in Mathematics 177.1, 2003, pp. 115–179
  • [HK06] André Henriques and Joel Kamnitzer “Crystals and coboundary categories” In Duke Mathematical Journal 132.2 Duke University Press, 2006, pp. 191–216 DOI: 10.1215/S0012-7094-06-13221-0
  • [Hum90] James E. Humphreys “Reflection Groups and Coxeter Groups”, Cambridge Studies in Advanced Mathematics Cambridge University Press, 1990
  • [Los19] Ivan Losev “Cacti and cells” In J. Eur. Math. Soc. 21.6, 2019, pp. 1729–1750
  • [Mos19] Jacob Mostovoy “The pure cactus group is residually nilpotent” In Archiv der Mathematik 113, 2019, pp. 229–235
  • [Tit69] Jacques Tits “Le problème des mots dans les groupes de Coxeter” In Symposia Mathematica Academic Press, 1969, pp. 175–185