• Aucun résultat trouvé

Generic Hecke algebra for Renner monoids

N/A
N/A
Protected

Academic year: 2021

Partager "Generic Hecke algebra for Renner monoids"

Copied!
26
0
0

Texte intégral

(1)

HAL Id: hal-00453780

https://hal.archives-ouvertes.fr/hal-00453780

Preprint submitted on 5 Feb 2010

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Generic Hecke algebra for Renner monoids

Eddy Godelle

To cite this version:

Eddy Godelle. Generic Hecke algebra for Renner monoids. 2010. �hal-00453780�

(2)

GODELLE EDDY

Abstract. We associate with every Renner monoidRageneric Hecke alge- bra H(R) overZ[q] which is a deformation of the monoidZ-algebra ofR. If M is a finite reductive monoid with Borel subgroupBand associated Renner monoidR, then we obtain the associated Iwahori-Hecke algebraH(M, B) by specialisingq inH(R) and tensoring by CoverZ, as in the classical case of finite algebraic groups. This answers positively to a long-standing question of L. Solomon.

2000 Mathematics Subject Classification: 20G40, 20C08, 20G05.

keywords : Renner monoid, Hecke algebra.

Introduction

Consider the groupG=GLn(Fq) of invertible matrices over the finite fieldFq. Denote by B its subgroup of upper triangular matrices, and by T its subgroup of diagonal matrices. Set ε = |1B|P

b∈Bb in C[G]. The quotient group NG(T)/T is isomorphic to the symmetric groupSn. Moreover, the Iwohori-HeckeC-algebra H(G,B) =εC[G]εis isomorphic to⊕w∈SnCwas aC-vector space, and the structure constants in the multiplicative table lie in Z[q]. More generally, if G is a finite reductive group over Fq, B is a Borel subgroup of G, and T a maximal torus included in B, then NG(T)/T is a Weyl group and the above results extend to the Hecke algebra H(G, B). Now, consider a finite reductive monoid M over Fq as defined by Renner in [21]. Such a monoid is a unit regular monoid and its unit group is a finite reductive group G. Solomon introduced in [23] the notion of a Iwahori-Hecke algebra H(M, B) of a finite reductive monoid M. Here, B is a Borel subgroup of G. This C-algebra is defined by H(M, B) = εC[M]ε where as beforeε= |B|1 P

b∈Bb in C[M]. In this framework, the Weyl group is replaced by an inverse monoidR, which is called the Renner monoid ofM. Its turns out that H(M, B) is isomorphic to ⊕r∈RCr as a C-vector space. An isomorphism is given byr7→T˜r=P

x∈BrBx. Therefore, this is natural to address the question of the existence of a normalisationTr =arr of the basis ( ˜Tr)r∈R such that in this new basis (Tr)r∈R, the structure constants in the multiplicative table lie in Z[q]

as in the case of finite reductive groups. Solomon considered this question in [23]

and answered in the positive in the specific case where M =Mn(Fq). In [24], he announced that in a forthcoming paper, he was going to extend his result and its proof to every finite reductive monoid that arises as the set of fixed points of a reductive monoid overFq (see Section 2.1 for a definition) by the Frobenius mapσ defined by σ(xi,j) = xqi,j. But it seems that this result has never be published.

In [19] Putcha proves that for every finite reductive monoid, one can normalised the basis ( ˜Tr)r∈Rsuch that the structure constants become rational inq. Howewer,

Date: February 5, 2010.

1

(3)

the question remained open, and Renner concluded in [22, sec. 8.3] that “the delicate part here is obtaining integral structure constants”. The main object of this article is to answer Solomon’s question in the positive for every finite reductive monoid.

We prove:

Theorem 0.1. Let M be a finite reductive monoid over Fq. Denote by R the as- sociated Renner monoid. There exists a normalisation of the basis ( ˜Tr)r∈R of the Iwahori-Hecke algebraH(M, B)such that the structure constants in the multiplica- tive table lie in Z[q]. Moreover, the coefficients of the polynomials depend on R only.

In Section 2, we provide explicit formulae (see Theorem 1.27), which are related to the existence of a length function onR. Moreover, we deduce a finite presentation ofH(M, B) in the spirit of the classical presentation ofH(G, B) (see Corollary 2.22 in Section 2.3).

Mokler, Renner and Putcha consider families of monoids that are closed to re- ductive monoids (see [13, 14, 15, 17, 18, 20] for instance. They are called finite monoids of Lie typeandface monoids. Indeed, finite reductive monoids are special cases of finite monoids of Lie type. To each of these groups can be associated a so-calledRenner monoid, whose properties are closed to Renner monoids of (finite) reductive monoids (See Examples 1.8 and 1.9 below). This explains why these monoids are still called Renner monoids in the latter references. However, there is some differences between these monoids (see Remark 1.10 for a discussion). We introduce here the notion of a generalised Renner monoid. All Renner monoids are examples of generalised Renner monoids. One motivation for this definition is to introduce a notion that plays for these various Renner monoids the role of the notion of a Coxeter system for Weyl group. We prove that all the properties shared by the various Renner monoids hold for generalised Renner monoid. In particular, it is a factorisable monoid and its unit group G is a Coxeter group. The crucial point regarding Solomon’s question is that we can associate with each such gener- alised Renner monoidR a generic Hecke algebra H(R) which is a ring on the free Z[q]-module with basisR. It turns out that Theorem 0.1 is a consequence of Theorem 0.2. LetM be a finite reductive monoid overFq with Renner monoidR.

The Iwahori-Hecke algebra H(M, B) is isomorphic to the C-algebra C⊗ZHq(R), whereHq(R)is the specialisation of the generic Hecke algebraH(R) atq.

The second main ingredient used in the proof of Theorem 0.1 is the existence of a length functionℓon every generalised Renner monoidR. This length function is related to the canonical generating setS∪Λ, which equips every generalised Renner monoid. In the case of reductive monoids, we investigate the relation of this length function with the product of double classes. We prove in particular that

Proposition 0.3. Let M be a reductive monoid with unit group G and Renner monoid R. Fix a maximal torusT and a Borel subgroupB that contains T in G.

(i) Letrlie inR ands lie inS, then

BsBrB =

BrB, if ℓ(sr) =ℓ(r);

BsrB, if ℓ(sr) =ℓ(r) + 1;

BsrB∪BrB, if ℓ(sr) =ℓ(r)−1.

(4)

(ii) Letr lie in R andelie inΛ, then then

BeBrB =BerB andBrBeB =BreB

This result extends results obtained in [5, 6], and leads to a similar result for finite reductive monoids.

The paper is organised as follows. In Section 1, we introduce the notion of a generalised Renner monoid, provide examples and investigate properties of such monoids. In particular, we define the length function ℓ and prove that a generic Hecke algebra can be associated with every generalised Renner monoid. In Sec- tion 2, we first recall the notion of a reductive monoid and prove Proposition 0.3.

Then we introduce the notion of a Iwahori-Hecke algebra in the context of monoid theory. We prove some motivating general results for such algebras. These results are probably well-known by semigroup experts, but we have not be able to find references for them. Finally, we turn to finite reductive monoids and conclude with the proof of Theorem 0.1 and 0.2.

1. Generic Hecke algebra

The notion of a Coxeter group has been introduced in order to study Weyl groups.

Our objective in this section is to develop a similar theory for Renner monoids. We need first to recall some standard notions and introduce useful notations.

1.1. Basic notions and notations. We refer to [8] for a general introduction on Semigroup Theory, and to [3] for a survey on factorisable inverse monoids. We refer to [1] for general theory and proofs on Coxeter systems.

1.1.1. Background on Semigroup Theory. IfM is a monoid, we letE(M) andG(M) its idempotent set and its unit group. We see a (lower)semi-lattice as a commuta- tive idempotent semigroup wherea≤biffab=ba=a. In particular,a∧b=ab. A semigroup isunit regular ifM =E(M)G(M) =G(M)E(M), and it isfactorisable if it is unit regular andE(M) is a semi-lattice. In this latter caseM is invertible, that is for every x in M there exists a unique y in M such that xyx = x (and thereforeyxy=y).

1.1.2. Background on Coxeter Group Theory.

Definition 1.1. Let Γ be a finite simple labelled graph whose labels are positive integers greater or equal than 3. We let denoteS the vertex set of Γ. We let E(Γ) denote the set of pairs ({s, t}, m) such that either{s, t} is an edge of Γ labelled bym, or{s, t}is not an edge of Γ andm= 2. When ({s, t}, m) belongs toE(Γ), we let|s, timdenote the wordsts· · · of lengthm. TheCoxeter groupW(Γ) associated with Γ is defined by the following group presentation

S

s2= 1 s∈S

|s, tim=|t, sim ({s, t}, m)∈ E(Γ)

In this case, one says that the pair (W(Γ), S) is aCoxeter system, and thatW is a Coxeter group. The Coxeter graph is uniquely defined by the Coxeter system.

Definition 1.2. Let (W, S) be aCoxeter system.

(i) Letwbelong toW. Thelength ℓ(w) ofwis the minimal integer ksuch thatw has a word representative of lengthk on the alphabetS. Such a word is called a

(5)

minimal word representative ofw.

(ii) The subgroup WI generated by a subsetI ofS is called a standard parabolic subgroup ofW.

A key tool in what follows is the following classical result.

Proposition 1.3. [1]Let (W, S)be a Coxeter systemwith Coxeter graph Γ.

(i) For every I⊆S, the pair (WI, I)is a Coxeter system. Its graph ΓI is the full subgraph of Γspanned byI.

(ii) For every I, J⊆S and every elementw∈W there exists a unique elementof minimal length in the double-class WJwWI. Furthermore there existsw1 inWI

andw2 inWJ such thatw=w2wwˆ 1 withℓ(w) =ℓ(w1) +ℓ( ˆw) +ℓ(w2).

Note that (ii) holds, in particular, whenIorJ are empty. The element ˆwis said to be (I, J)-reduced. In the sequel, we let Red(I, J) denote the set of (I, J)-reduced elements. Note also that the pair (w1, w2) is not unique in general, but it becomes unique if we require thatw2wˆ is (∅, J)-reduced (or that ˆww1is (I,∅)-reduced).

1.2. Generalised Renner monoids.

1.2.1. Generalised Renner-Coxeter System. If R is a factorisable monoid and e belongs toE(R) we letW(e) andW(e) denote the subgroups defined by

W(e) ={w∈G(R)|we=ew}

W(e) ={w∈G(R)|we=ew=e}.

The unit groupG(R) acts onE(R) by conjugacy.

Definition 1.4. (i) Angeneralised Renner-Coxeter systemis a triple (R,Λ, S) such that

(ECS1) Ris a factorisable monoid;

(ECS2) Λ is both a transversal of E(R) for the action of G(R) and a sub-semi- lattice;

(ECS3) (G(R), S) is a Coxeter system;

(ECS4) for every pair e1 ≤ e2 in E(R) there exists w in G(R) and f1 ≤f2 in Λ such thatwfiw−1=eifori= 1,2;

(ECS5) for every e in Λ, the subgroups W(e) and W(e) are standard Coxeter subgroups ofG(R);

(ECS6) the map e ∈ Λ 7→ λ(e) = {s ∈ S | se = es 6= e} is not decreasing:

e≤f =⇒ λ(e)⊆λ(f).

In this case, we say thatRis ageneralised Renner monoid. Following the standard terminology for Renner monoids, we call the section Λ thecross section latticeofR, and we define thetype mapofRto be the mapλ: Λ→Sdefined byW(e) =Wλ(e). Notation 1.5. fore inΛ, we set

λ(e) ={s∈S|se=es=e}

W(e) =Wλ(e)

Remark 1.6. Assume (R,Λ, S) is a generalised Renner-Coxeter system.

(i) SinceW(e) is a standard Coxeter subgroup of W(e), we have W(e) =Wλ(e).

(6)

Moreover, This is clear thatW(e) is a normal subgroup ofW(e). As a consequence, W(e) =W(e)×W(e) andλ(e) =λ(e)∪λ(e).

(ii) Below, several results can be proved without assuming Property (ECS6). How- ever this is a crucial tool in the proof of Theorem 1.27 and Proposition 1.17.

(iii) IfE(R) is finite and a lower semi-lattice, then it has to be a lattice. This is so forRenner monoids associated withreductive monoids.

(iv) the mapλ is not increasing:

e≤f =⇒ λ(f)⊆λ(e).

(v) We can haveλ(e) =λ(f) andλ(e) =λ(f) fore6=f (see [6, Sec. 2.3]).

Now we provide some examples of generalised Renner monoids.

Example 1.7. LetM be a reductive monoid (see Section 2.1 for a definition, and Example 2.6). The associated Renner monoidR(M) ofM is a generalised Renner monoid by [22].

Example 1.8. LetM be aabstract finite monoid of Lie type (see [17],[20] or [22]

for a definition. Note that these groups are calledregular split monoidsin [17], and finite monoids of Lie type in [20]). The associatedRenner monoid R(M) ofM is a generalised Renner monoid. Property (ECS6) follows from [17, Cor. 3.5(i)]. The other defining properties hold by [18, Sec. 2]. The seminal examples of an abstract finite monoid of Lie type is a Renner monoid of afinite reductive monoid [21]. In Section 3 we focus on these monoids.

Example 1.9. LetGbe a Kac-Moody group over a field Fof characteristic zero whose derived group is the special Kac-moody group introduced in [11, 12]. De- note by (W, S) the associated Coxeter system. The Coxeter group W is infinite.

LetF a(X) be the set offacesof its associated Tits coneX(see [13] for details). The action ofW onX induces an action on the latticeF a(X). TheRenner monoid R is the monoidW⋉F a(X)/∼where∼is the congruence onW⋉F a(X) defined by (w, R)∼(w, R) if R=R and w′−1wfixesR pointwise [13]. ThenR is a gener- alised Renner monoid. Properties (ECS1), (ECS2), (ECS3) and (ECS5) are proved in [13] (see also [15]). The cross section lattice Λ can be identified with the set of infinite standard parabolic subgroups ofW that have no finite proper normal stan- dard parabolic subgroups. The semi-lattice structure is given byWI ≤WJifJ⊆I.

If Θ belongs to Λ, then λ(Θ) = Θ and λ(Θ) ={s∈ S | ∀t ∈Θ, st=ts}. The latter equality clearly implies (ECS6). Finally, Property (ECS4) can be deduced from [15, Theorem 2 and 4].

Remark 1.10. In Examples 1.7, 1.8 and 1.9 we provide examples of generalised Renner monoids that are all called Renner monoid in the literature. From our point of view, this is not a suitable terminology since there is crucial differences between these monoids. Therefore, using the same terminology may be misleading.

For instance, for Renner monoids of reductive monoids one hasλ(e) =T

f≤eλ(f) andλ(e) =T

f≥eλ(f). This is not true in general for Renner monoids associated with abstract finite monoids of Lie type (see [20] for a details). In Renner monoids of reductive monoids and of abstract finite monoids of Lie type, all maximal chains of idempotents have the same size. This is not true for Renner monoids of example 1.9, as explained in [13].

(7)

1.2.2. Presentation for generalised Renner monoids. For all this section, we fix a generalised Renner-Coxeter system (R,Λ, S). We letW denote the unit group ofR.

Our objective is to prove that important properties shared by Renner monoids of Examples 1.7, 1.8, 1.9 can be deduced from their generalised Renner-Coxeter system structure. In particular, we extend to this context the results obtained in [6]. By Proposition 1.3, For every w in W and every e, f in Λ, each of the sets wW(e), W(e)w, wW(e), W(e)w and W(e)wW(f) has a unique element of minimal length. In order to simplify notation, we set Red(·, e) = Red(∅;λ(e)), Red(e,·) = Red(λ(e),∅); Red(·, e) = Red(∅, λ(e)); Red(e,·) = Red(λ(e),∅);

Red(e, f) = Red(λ(e), λ(f)).

Proposition 1.11. For everyr inR,

(i) there exists a unique triple (w1, e, w2) with e ∈ Λ, w1Red(·, e) and w2Red(e,·) such thatr=w1ew2;

(ii) there exists a unique triple (v1, e, v2) with e ∈ Λ, w1Red(·, e) and w2Red(e,·)such that r=v1ev2

Following [22], we call the triple (w1, e, w2) the normal decomposition ofr.

Proof. Let us prove (i). The proof of (ii) is similar. Letrbelong to the monoidR.

By Property (ECS1), there exists e in E(R) and w in W such that r = ew. By Property (ECS2) there exists e1 in Λ and v in W such that e = ve1v−1. Then r =vew1 with w1 =v−1w. By Remark 1.6(i), we can write v = v1v1 and w1 = w2w′′2w2 with v1, w2, v1, w2 and w′′2 in Red(·, e), Red(e,·), W(e), W(e) and W(e), respectively. Then we haver=v1w2ew2, and v1w2 belongs to Red(·, e), still by Remark 1.6(i). Now assume r = w1ew2 = v1f v2 with e, f in Λ, w1, v1

in Red(·, e) and Red(·, f), respectively, and w2, v2 in Red(e,·) and in Red(f,·), respectively. Then (w1w2)w2−1ew2= (v1v2)v−12 f v2. This impliesw2−1ew2=v−12 f v2

by [3]. As a consequence, e =f and v2w−12 lies in W(e). Since v2 and w2 both belong to Red(e,·), we must have v2 =w2. Now, it follows that w1e = v1e and w−11 v1 lies inW(e). This implies w1=v1 in Red(·, e).

Lemma 1.12. Let e, f belong toΛ andw lie in Red(e, f).

(i) There existshinΛ such thatw belongs toW(h)andewf =wh.

(ii) The element wlies inW(h). Therefore, wh=h.

Note that in the above lemma we have h ≤ e∧f = ef. In the sequel the elementhis denoted bye∧wf.

Proof. The proof is similar to [6, Prop 1.21]. (i) Consider the normal decompo- sition (w1, h, w2) of ewf. By definition w1 belongs to Red(·, h) and w2 belongs to Red(h,·). The element w−1ewf is equal to w−1w1hw2 and belongs to E(R).

Since w2 lies in Red(h,·), this implies that w3 = w2w−1w1 lies in W(h), and that f ≥ w−12 hw2. By Property (ECS4), there exists w4 in W and f1, h1 in Λ, with f1 ≥ h1, such that w−14 f1w4 = f and w4−1h1w4 = w−12 hw2. Since Λ is a cross section for the action of W, we have f1 = f and h1 = h. In particular, w4 belongs to W(f). Since w2 belongs to Red(h,·), we deduce that there ex- ists r in W(h) such that w4 = rw2 with ℓ(w4) = ℓ(w2) +ℓ(r). Then w2 lies in W(f), too. Now, write w1 = w1w1′′ where w1′′ lies in W(h) and w1 belongs to Red(·, h). One has ewf =w1hw1′′w2, and w1w′′2 lies in Red(h,·). By symmetry, we get thatw1 belongs to W(e). The elementw′−11ww2−1 is equal tow′′1w−13 and

(8)

belongs to W(h). But, by hypothesis w lies in Red(e, f). Then we must have ℓ(w′′1w3−1) = ℓ(w′−11) +ℓ(w) +ℓ(w−12 ). Since w1′′w−13 belongs to W(h), it follows that w1 andw2 belong to W(h) too. This implies w2 =w1 = 1 andw =w′′1w3−1. Therefore,ewf =hw1′′=hw=wh.

(ii) This is a direct consequence of the following fact: forh, ein Λ such thath≤e, we haveW(h)∩Red(e,·)⊆W(h) and W(h)∩Red(·, e)⊆W(h). Assumewlies in W(h)∩Red(·, e), then we can writew=w1w2=w2w1 wherew1 lies inW(h) and w2 lies inW(h). Sinceh≤e, we have λ(h)⊆λ(e) and W(h)⊆W(e).

Sincewbelongs to Red(·, e), this impliesw2= 1. The proof of the second inclusion is similar.

Corollary 1.13. (i) For every chain e1 ≤e2≤ · · · ≤em in E(R) there exists w in G(R) and a chain f1 ≤ f2 ≤ · · · ≤ fm in Λ such that wfiw−1 =ei for every index i.

(ii) IfΛ has an infimume, thenλ(e) =S.

(iii) For all e, f in Λand win Red(e, f), one has

ewf = max{h∈Λ|h≤e, h≤f, w∈W(h)}=f w−1e.

In the case of Renner monoids of reductive monoids, the lattice Λ has an infi- mumeandλ(e) =λ(e) =S. In other words,eis a zero element ofR.

Proof. (i) Assume w1e1w1−1 ≤ · · · ≤wmemwm−1. We prove the result by induction onm. For m= 2 this is true by Property (ECS4). Assumem≥3. By induction hypothesis, we can assumew2=· · ·=wm. We can also also assume thatw1belongs to Red(·, e1). By hypothesis, we havew1e1w−11 w2e2w−12 =w1e1w−11 . We can write w−11 w2 = v1v3v−12 with v1 in W(e1), v2 in W(e2) and v3 in Red(e1, e2). Then w1e1w−11 w2e2w−12 =w1v1e1v3e2v2−1w−12 . Ifv36= 1, then we get a contradiction by Lemma 1.12(i) and Proposition 1.11. Then v2 = 1 ande1e2=e1. It follows that w1v1=w2v2. Writev1=v1⋆v1andv2=v2⋆v2withvi⋆inW(ei) andviinW(ei).

We havew1v1⋆v−12⋆ =w2v2v1−1. Sinceλ(e2)⊆λ(e1) andλ(e1)⊆λ(e2), we get that v1⋆v−12⋆ and v2v1−1 lie in W(e1) andW(e2), respectively. Then w1e1w−11 = we1w−1 and w2e2w2−1 = we2w−1 with w = w1v1⋆v−12⋆. But W(e2) ⊆ W(ej) for j∈ {2,· · ·, m}. Therefore,w2ejw2−1=wejw−1for everyj≥2.

(ii) ifs∈S does not belong toλ(e), thenese < ein Λ.

(iii) This is clear that e∧wf lies in {h∈ Λ | h ≤e, h ≤ f, w ∈ W(h)}. Now, if h∈ Λ verifies h≤ e, h ≤f, and w ∈ W(h), then h(e∧wf) = hw−1(ewf) = w−1hwf =hf =h. Therefore,h≤ewf. The last equality follows form the fact

thatw−1 belongs to Red(f, e).

Proposition 1.14. For everywin W, we fix an arbitrary reduced word represen- tative w. We set Λ = Λ\ {1}. The monoid R admits the monoid presentation whose generating set isS∪Λ and whose defining relations are:

(COX1) s2= 1, s∈S;

(COX2) |s, tim=|t, sim, ({s, t}, m)∈ E(Γ);

(REN1) se=es, e∈Λ,s∈λ(e);

(REN2) se=es=e, e∈Λ,s∈λ(e);

(REN3) ewf =e∧wf, e, f ∈Λ,w∈Red(e, f).

(9)

Proof. This is clear that the relations stated in the proposition hold in R. Con- versely, every elementrinRhas a unique representing wordwevsuch that (w, e, v) is its normal decomposition, and this is immediate that every representing word ofr onS∪Λ can be transformed intowevusing the given relations only.

Remark 1.15. (i) The above presentation is not minimal in general. Some of the relations of type (REN3) can be removed (see the proof of [6, Theorem 0.1] and Remark 1.32 below).

(ii) The reader may verify that the result of Proposition 1.14 and its proof still hold if we do not assume Property (ECS6), except that Relation (REN3) must be replace by

(REN3’) ewf =w(e∧wf), e, f∈Λ,w∈Red(e, f).

Indeed, Lemma 1.12(i) still hold.

One may wonder whether every monoid defined by a monoid presentation like in Proposition 1.14. The answer is positive under some necessary assumptions:

Definition 1.16. Ageneralised Renner-Coxeter data is 4-uple (Γ,Λ, λ, λ) such that Γ is a Coxeter graph with vertex setS, Λ is a lower semi-lattice andλ

are two maps from Λ toS that verifies

(a) for every e in Λ, the graphs spanned by λ(e) and λ(e) in Γ are not connected, and

e≤f ⇒λ(f)⊆λ(e) andλ(e)⊆λ(f).

(b) for everyf, g in Λ and everyw∈Red(f, g) the set e∈Λ|e≤f, e≤g andw∈Wλ(e)

has a greatest element, denoted byf ∧wg.

withλ(e) =λ(e)∪λ(e) fore∈Λand Red(e, f) = Red(λ(e), λ(f)) in the Coxeter groupW(Γ) associated with Γ.

Note that properties (a) and (b) hold in every generalised Renner-Coxeter sys- tem. Actually, if Λis any lower semi-lattice such that all maximal chains are finite, then Assumption (b) is necessarily verified.

Theorem 1.17. Assume M is a monoid. There exists a generalised Renner- Coxeter system (M,Λ, S) if and only if there exists a generalised Renner-Coxeter data(Γ,Λ, λ, λ), whereS is the vertex set ofΓ, such thatM admits the following monoid presentation

(COX1) s2= 1, s∈S;

(COX2) |s, tim=|t, sim, ({s, t}, m)∈ E(Γ);

(REN1) se=es, e∈Λ,s∈λ(e);

(REN2) se=es=e, e∈Λ,s∈λ(e);

(REN3) ewf =e∧wf, e, f ∈Λ,w∈Red(e, f).

Wherew is an arbitrary fixed minimal representing word ofw∈W(Γ).

In this case,W(Γ)is canonically isomorphic to the unit group ofM, andΛ embeds inM withΛ = Λ∪ {1}.

Note that given a generalised Renner-Coxeter data (Γ,Λ, λ, λ), Relations (COX1) and (COX2) implies that the monoidM defined by the presentation stated in Theorem 1.17 does not depend on the chosen representing wordsw. Theorem 1.17 follows from the following lemmas.

(10)

Lemma 1.18. Consider a generalised Renner-Coxeter data (Γ,Λ, λ, λ)and the monoidM defined by the presentation stated in Theorem 1.17. Then for every f, g inΛ and every w∈Red(f, g),

(b1) e∧1f =e∧f ande∧wf ≤e∧f; (b2) e∧wf =f∧w−1e;

(b3) w∈Wλ(e∧wf).

Proof. Properties (b1) and (b2) are immediate consequences of Assumption (b).

Properties (b3) follows from Assumption (a). The main argument is like in the proof of Lemma 1.12(ii). If w doesnot belong to Wλ(e∧wf), then we can write w =ww with w ∈ Wλ(e∧wf) and w ∈Wλ(e∧wf). ButWλ(e∧wf) ⊆Wλ(f)

andwlies in Red(e, f). Therefore,w= 1.

Lemma 1.19. Consider a generalised Renner-Coxeter data (Γ,Λ, λ, λ)and the monoid M defined by the presentation stated in Theorem 1.17. Let FM(S∪Λ)be the free monoid on S∪Λ, andbe the congruence on FM(S∪Λ)generated by the defining relations of M. Hence by definition, M is equal to FM(S∪Λ)/≡.

(i) If ω1 and ω2 are two words on S such that ω1 ≡ ω2, then they represent the same element in W(Γ).

(ii) If e lie in Λ and ω lie in FM(S∪Λ) with e ≡ ω, then the word ω is equal to ν1e1ν2· · ·ekνk+1 where for every i we have e ≤ ei in Λ and νi are words on S whose images in W(Γ) belong to Wλ(e). Furthermore, the image of the wordν1ν2· · ·νk+1 inWλ(e)=Wλ(e)/Wλ(e)is trivial.

Proof. In this proof we writeω1=ω˙ 2if the two wordsω12are equals. If the words ω1ω2represent the elementsw1, w2inM, respectively, thenω1=ω˙ 2impliesω1≡ω2

and w1 =w2. Conversely,w1 =w2 if and only if ω1 ≡ω2. Point (i) is clear: if ω1 ≡ω2 then one can transformω1 into ω2 using relations (COX1) and (COX2) only, since the words in both sides of Relations (REN1-3) contain letters in Λ. Let us prove (ii). Write ω11 ω2 if one can transform ω1 into ω2 by applying one defining relation ofM onω1. Ife≡ω, then there existsω0=e, ω˙ 1,· · ·, ωr=ω˙ such thatωi1ωi+1. We prove the result by induction onr. If r= 0 we have nothing to prove. Assumer≥1. By induction hypothesis,ωr−1=µ˙ 1f1µ2· · ·µjfjµj+1 with e≤fi in Λ andµi is a word onS whose image inW(Γ) belongs toWλ(e), and the image of the wordµ1µ2· · ·µj+1inWλ(e)=Wλ(e)/Wλ(e)is trivial. We deduce the result forω=ω˙ rby considering case by case the type of the defining relation applied toωr−1 to obtainωr. The cases where the relation is of one of the types (COX1), (COX2) or (REN1) are trivial. The case where the relation is of type (REN2) follows from Property (a) in Definition 1.16: by induction hypothesis, one has λ(fi)⊆λ(e)⊆λ(e). Finally, the case where the relation is of type (REN3) follows from properties (a) and (b) by Lemma 1.18. If the imageui ofµiinW(Γ) belongs to Red(fi−1, fi) with µi = ui and ω=µ˙ 1f1· · ·µi−1(fiuifi+1i+1fi+2· · ·fjµj+1

then e≤fi−1uifi. Conversely, if ω=µ1f1µ2· · ·fi−1µieiuiei+1µi+1· · ·µjfjµj+1

wherefi =eiuiei+1forei, ei+1in Λand someuiin Red(ei, ei+1), thene≤fi≤ei

ande≤fi≤ei+1; Moreover,uibelongs toWλ(fi), which is included inWλ(e). In all these cases the wordsν1ν2· · ·νk+1andµ1µ2· · ·µj+1represent the same element inWλ(e)/Wλ(e), which is trivial by induction hypothesis.

Proof of Theorem 1.17. Consider a generalised Renner-Coxeter system (M,Λ, S).

Denote by Γ the Coxeter graph with vertex setS of the unit group ofM, and set

(11)

Λ = Λ\ {1}. It follows from previous results that (Γ,Λ, λ, λ) is a generalised Renner-Coxeter data, and by Proposition 1.14 that M has the required monoid presentation. Conversely, consider a generalised Renner-Coxeter data (Γ,Λ, λ, λ) and letM denote the monoid defined by the presentation stated in Theorem 1.17.

By Lemma 1.19(i), the subgroup ofM generated byScan be identified withW(Γ).

Lemma 1.19(ii) implies that Λ injects in M, as a set. Let e, f be in Λ. In M one has ef = f e = e∧1f = e∧f. Assume furthermore that w lies in W(Γ).

Lemma 1.19(ii) implies also that (wew−1)f =wew−1 if and only ife≤f in Λ and wlie inWλ(e). Letwew−1 andvf v−1 be inE(M) withe, f in Λ. Writew−1v= v1v2v3 withv2in Red(e, f),v1inWλ(e) andv3in Wλ(f). Thenev2f =e∧v2f and v2 lies inWλ(e∧v2f). We get,

wew−1vf v−1=wv1e∧v2f v3v−1=wv1f∧−1v2 ev3v−1=wv1v2f v−12 ev2v3v−1= wv1v2v3f v3−1v2−1v1−1ev1v2v3v−1=vf v−1wew−1. It is easy to see that every representing wordωonS∪Λof en elementwofM can be transformed into a wordω12≡ω1−11 ω1ω2whereebelongs to Λ = Λ∪ {1}

and ω1, ω2 represent words in W(Γ). Moreover, if ω contains some letter in Λ, then e has to be in Λ. Therefore, M is unit regular and G(M) = W(Γ). In particular Property (ECS3) holds. Assume w=w1ew2 lies inE(M) with w1, w2

in W(Γ) and e in Λ. Ife= 1 then w1w2 has to be equal to 1 in W(Γ). Assume e6= 1. Then w1ew2w1ew2 =w1ew2, and ew2w1e=e. By Lemma 1.19(ii), w2w1

belongs to Wλ(e) and w=w1ew1−1. ThusE(M) ={wew−1 |e ∈Λ, w ∈W(Γ)}

is a semi-lattice and Property (ECS1) holds. Let w1, w2, v1, v2 be in W(Γ) and e, f be in Λ such that w1ew2 =v1f v2 in M. Then e=w−11 v1f v2w−12 ande≤f. By symmetry, e = f and the elements w−11 v1 and v2w2−1 belong to Wλ(e). This implies that Λ is a transversal of E(M) for the action of W(Γ) and a sub-semi- lattice of E(M). Therefore, we get Property (ECS2). Furthermore, if w2 =v1 = 1 and v2 = w1, then w1 lies in Wλ(e). If w2 = v1 = v2 = 1, then w1 lies in Wλ(e) by Lemma 1.19(ii). Property (ECS5) follows. If wew−1 ≤ vf v−1, then wew−1vf v−1=wew−1 and ew−1vf v−1w=eThen w−1v lies inWλ(e)×Wλ(e), which is included in Wλ(e)×Wλ(f). As a consequence, Property (ECS4) holds.

Finally, Property (ECS6) holds by hypothesis.

1.2.3. Length function for generalised Renner-Coxeter systems. As explained in the introduction, to answer Solomon’s question, we need to define a length function on finite reductive monoids. Here we introduce this length function in the general context of generalised Renner-Coxeter systems. This extends results obtained in [5]

and [6]. As before, (R,Λ, S) is a generalised Renner-Coxeter system. The unit group ofR is denoted byW, and we set Λ= Λ\ {1}.

Definition 1.20. (i) We setℓ(s) = 1 forsinSandℓ(e) = 0 forein Λ. Letx1, . . . , xk

be in S∪Λ and consider the wordω=x1· · ·xk. Then the length of the wordω is the integerℓ(ω) defined byℓ(ω) =Pk

i=1ℓ(xi).

(ii) Thelengthof an elementwwhich belongs toRis the integer ℓ(w) defined by ℓ(w) = min{ℓ(ω)|ω is a word representative ofwover S∪Λ}.

Ifωis a word representative ofωsuch thatℓ(w) =ℓ(ω), we say thatω is aminimal word representative ofw.

(12)

Proposition 1.21. Let rbelong toR.

(i) The length functiononR extends the length functiondefined onW. (ii) ℓ(r) = 0iff rlies inΛ.

(iii) Ifslies inS then|ℓ(sr)−ℓ(r)| ≤1.

(iv) If r belongs to R, thenℓ(rr)≤ℓ(r) +ℓ(r).

Proof. This is direct consequences of the definition of the length function.

Proposition 1.22. Let r belong to R. If (w1, e, w2) is the normal decomposition of r, then

ℓ(r) =ℓ(w1) +ℓ(w2).

Proof. Using the relations of the monoid presentation of R stated in Proposi- tion 1.14, every representative word of r can be transformed into w1ew2 without increasing the length. Thereforeℓ(r) =ℓ(w1) +ℓ(e) +ℓ(w2) =ℓ(w1) +ℓ(w2).

From the proof of the above proposition, we also deduce that

Corollary 1.23. Letrbelong toRandω1, ω2be two minimal word representatives ofr. Using the relations of the monoid presentation ofRstated in Proposition 1.14, one can transformω1 intoω2 without increasing the length.

1.2.4. Matsumoto’s Lemma for generalised Renner-Coxeter systems. In this section we state and prove some technical results that play the role ofMatsumoto’s Lemma in the context of generalised Renner-Coxeter systems. we need these results when proving Theorem 1.27. As before, (R,Λ, S) is a generalised Renner-Coxeter system.

Let us first recall Matsumoto’s Lemma.

Lemma 1.24. [9, Sec. 7.2]Consider a Coxeter system (W, S). Letwbelong to W ands, tbelong toS. If ℓ(swt) =ℓ(w)andℓ(sw) =ℓ(wt), then sw=wt.

Lemma 1.25. Letrbelong toRands, tbelong toS. Let(w1, e, w2)be the normal decomposition ofr. Then

(i)ℓ(sr) =ℓ(r)±1 if and only if the normal decomposition ofsris(sw1, e, w2). In this case,ℓ(sr)−ℓ(r) =ℓ(sw1)−ℓ(w1).

(ii)ℓ(sr) =ℓ(r)if and only ifsr=rif and only ifsw1=w1ufor someuinλ(e).

In this case, ℓ(sw1) =ℓ(w1) + 1.

(iii) ℓ(rt) = ℓ(r)±1 if and only if the normal decomposition of rt is either (w1, e, w2t) or (w1u, e, w2) for some u in λ(e). Furthermore, in the former case ℓ(rt)−ℓ(r) = ℓ(w2t)−ℓ(w2), and in the latter case w2t = uw2 with ℓ(w2t) = ℓ(w2) + 1.

(iv)ℓ(rt) =ℓ(r)if and only ifr=rt if and only ifw2t=uw2 for some uinλ(e).

(v) Ifℓ(srt) =ℓ(r)andℓ(sr) =ℓ(rt)6=ℓ(r), then there exists uinλ(e)such that sw1=w1uanduw2=w2t. As a consequence, sr=rt.

Proof. Recall that |ℓ(sr)−ℓ(r)| ≤ 1 and |ℓ(rt)−ℓ(r)| ≤ 1. The normal de- composition of sr is (sw1, e, w2) if and only if sw1 belongs to Red(·, e). Since w1 belongs to Red(·, e), this is clearly the case if ℓ(sw1) = ℓ(w1)−1. As- sume ℓ(sw1) = ℓ(w1) + 1 and sw1 does not belong to Red(·, e). Then we can write sw1 =w1ufor someuin λ(e) such thatℓ(sw1) =ℓ(w1) + 1. In particular, ℓ(sw1u) =ℓ(w1) =ℓ(w1). On the other hand,ℓ(w1u) =ℓ(w1) + 1 =ℓ(sw1) because w1 belongs to Red(·, e), andulies inλ(e). By Lemma 1.24, we get sw1 =w1u andsr=sw1ew2=w1uew2=w1ew2=r. This proves (i) and (ii) since the other

(13)

implications are obvious. The normal decomposition ofrtis (w1, e, w2t) if and only ifw2t belongs to Red(e,·). Sincew2 belongs to Red(e,·), this is clearly the case if ℓ(w2t) =ℓ(w2)−1. Assumeℓ(w2t) =ℓ(w2)+1 andw2tdoes not belong to Red(e,·).

Then we can writew2t=uw2 for someuin λ(e) such thatℓ(w2t) =ℓ(w2) + 1. As before we can conclude thatw2t=uw2. Ifulies inλ(e) thenrt=r. Otherwise,u belongs toλ(e) andw1ubelongs to Red(·, e). This is true sinceubelongs toλ(e) and therefore commutes with each element of λ(e). Then the normal decompo- sition ofrt is (w1u, e, w2). This proves (iii) and (iv). Now assume ℓ(srt) = ℓ(r) and ℓ(sr) = ℓ(rt) 6= ℓ(r). We claim that ℓ(w2t) = ℓ(w2) + 1 and there exists u in λ(e) such that uw2 = w2t. If it was not the case, by above arguments, the normal decomposition of srt would be (sw1, e, w2t) and ℓ(srt) = ℓ(r)±2. Since we assume ℓ(rt)6=ℓ(r), the elementuhas to belong to λ(e). Finally, using that ℓ(sr) =ℓ(rt)6=ℓ(r) =ℓ(srt) we deduce thatℓ(sw1) =ℓ(w1u) andℓ(w1) =ℓ(sw1u),

which in turn impliessw1=w1uby Lemma 1.24.

Lemma 1.26. Let rbelong toR,sbelong toS andf belong toΛ. Let(w1, e, w2) be the normal decomposition of r.

(i) If ℓ(rf) =ℓ(r)thenw2 belongs toW(f).

(ii) Ifℓ(f r) =ℓ(r) thenw1=w1w1′′ wherew1 lies inW(f)andw1′′ lies inW(e).

(iii) Ifℓ(sr) =ℓ(r)−1, thenℓ(srf)≤ℓ(rf). Ifℓ(sr) =ℓ(r)+1, thenℓ(srf)≥ℓ(rf).

(iv) Ifℓ(rs) =ℓ(r)−1, thenℓ(f rs)≤ℓ(f r). Ifℓ(rs) =ℓ(r)+1, thenℓ(f rs)≥ℓ(f r).

Proof. By definition of the normal decomposition, w2 belongs to Red(e,·). Write w2=w2w2′′withw2, w′′2 in the unit groupW ofRsuch thatℓ(w2) =ℓ(w2) +ℓ(w′′2), w′′2 belongs toW(f) and w2 belongs to Red(·, f). Then w2 lies in Red(e, f). By Relation (REN3), we have rf = w1(e∧w2 f)w′′2. It follows that ℓ(w2) = 0, and w2=w′′2. This proves (i). The prove of (ii) is similar except that we need first to decomposew1 inw1w′′1 wherew′′1 lies inW(e) andw1lies in Red(·, e).

(iii) Assume ℓ(sr) =ℓ(r)−1. Writew1 =sv1 with ℓ(w1) =ℓ(v1) + 1, and write w2 = w2w′′2v2′′′ with w2, w′′2, w2′′′ in W such that ℓ(w2) = ℓ(w2) +ℓ(w2′′) +ℓ(w′′′2), where w′′2 belongs to W(f), w2′′′ belongs to W(f) andw2 belongs to Red(e, f).

Then (v1, e, w2) is the normal decomposition of sr. One has srf = v1ew2f w′′2 = v1ew′′2 where e = e∧w2 f belongs to Λ. Writew′′2 = v′′2v2v2 such that ℓ(w′′2) = ℓ(v2′′) +ℓ(v2) +ℓ(v2) with v2′′ ∈ W(e), v2 ∈ W(e) and v2 ∈ Red(e,·). We claim that v2′′ = 1. Indeed w2 belongs to W(e) by Lemma 1.12(ii), and w2 = w2v′′2v2v2w2′′′ =v2w2v′′2v2w2′′′ with ℓ(w2) =ℓ(v2) +ℓ(v2) +ℓ(w2) +ℓ(v′′2) +ℓ(w′′′2).

Butv2∈W(e)⊆W(e), since e ≤eby Property (ECS6), whereasw2 belongs to Red(e,·) by definition of the normal decomposition. Hence, v2 = 1. Now, write v1 = v1v1′′ such that ℓ(v1) = ℓ(v1) +ℓ(v1′′) with v1 ∈ Red(·, e) and v1′′ ∈ W(e). Then srf = v1ev2 and (v1, e, v2) is the normal decomposition of swf. Sinceℓ(ssr) =ℓ(sr) + 1, we have ℓ(sv1v′′1) =ℓ(sv1) =ℓ(v1) + 1 by Lemma 1.25(i).

This implies ℓ(sv1) =ℓ(v1) + 1 and we cannot haveℓ(ssrf) = ℓ(srf)−1, still by Lemma 1.25(i). Assumeℓ(sr) =ℓ(r)+1. let (v1, e, w2) be the normal decomposition of r, and (v1, e, v2) be the normal decomposition of rf. It follows from above arguments that v1 left dividesv1. We conclude using Lemma 1.25: ℓ(sr) =ℓ(r) + 1⇒ℓ(sv1) =ℓ(v1) + 1⇒ℓ(sv1) =ℓ(v1) + 1⇒ℓ(srf)≥ℓ(rf). The proof of (iv) is similar.

(14)

1.3. Free module overR. For all this section, we assume (R,Λ, S) is a generalised Renner-Coxeter system. We letW denote the unit group ofR, and set Λ = Λ\{1}.

We fix an arbitrary unitary associative ringA. We letV denote the freeA-module with basis elementsTrforr∈R.

Theorem 1.27. FixqinA. There exists a unique structure of unitary associative A-algebra onV such thatT1 is the unity element and the following conditions hold for everyxin S∪Λ and every rin R:

TxTr=Txr, ifx∈S andℓ(xr) =ℓ(r) + 1;

TxTr=qTr, ifx∈S andℓ(xr) =ℓ(r);

TxTr= (q−1)Tr+qTxr, ifx∈S andℓ(xr) =ℓ(r)−1;

TxTr=qℓ(r)−ℓ(xr)Txr, ifx∈Λ.

We follow the method explained in [10, Sec. 7.1] for the Hecke algebra of Coxeter groups. Let E =EndA(V) the A-algebra of endomorphisms of the A-module V. ForsinS andrin R, we define ρs inE by

ρs(Tr) =Tsr, ifℓ(sr) =ℓ(r) + 1;

ρs(Tr) =qTr, ifℓ(sr) =ℓ(r);

ρs(Tr) = (q−1)Tr+qTsr, ifℓ(sr) =ℓ(r)−1.

Forein Λ andrinR, we defineρeby

ρe(Tr) =qℓ(r)−ℓ(er)Ter

Similarly, forsin S andrinR, we defineρsin E by

ρs(Tr) =Trs, ifℓ(sr) =ℓ(r) + 1;

ρs(Tr) =qTr, ifℓ(r) =ℓ(rs);

ρs(Tr) = (q−1)Tr+qTrs, ifℓ(sr) =ℓ(r)−1.

Forein Λ andrinR, we defineρeby

ρe(Tr) =qℓ(r)−ℓ(re)Tre.

The key tool in the proof of Theorem 1.27 is the following result.

Lemma 1.28. For everyx, y inS∪Λ,

ρxρyyρx.

Proof. Letr belong toR and x, y belong toS∪Λ. We prove that ρxy(Tr)) = ρyx(Tr)). Clearly we can assume x 6= 1 and y 6= 1. By Proposition 1.21, ℓ(xry)≤ℓ(x) +ℓ(r) +ℓ(y)≤(r) + 2. We provide case by case as in [9].

Case 1: ℓ(xry) =ℓ(r) +ℓ(x) +ℓ(y). We must have ℓ(xr) = ℓ(r) +ℓ(x), ℓ(ry) = ℓ(r) +ℓ(y) and ℓ(xry) = ℓ(ry) +ℓ(x) = ℓ(xr) +ℓ(y). Therefore ρxy(Tr)) = ρx(Try) =Txryy(Txr) =ρyx(Tr)).

Case 2: ℓ(xry) =ℓ(r) + 1. We must have ℓ(xr) ≥ℓ(r), ℓ(ry) ≥ℓ(r), and x or y, possibly both, belongs toS. Ifxorybelongs to Λ, we are in Case 1. So we assume xandy belong toS.

Subcase 1: ℓ(xr) = ℓ(r), that is xr = r. Then ℓ(ry) = ℓ(xry) = ℓ(r) + 1 and ℓ(xry) = ℓ(xr) + 1. Therefore ρxy(Tr)) = ρx(Try) = qTxry = ρy(qTxr) = ρyx(Tr)). The caseℓ(ry) =ℓ(r) is similar.

Subcase 2: ℓ(ry) = ℓ(xr) = ℓ(r) + 1. Then ℓ(ry) = ℓ(xr) = ℓ(xry). We deduce thatρxy(Tr)) =ρx(Try) =qTxryy(Txr) =ρyx(Tr)).

Références

Documents relatifs

We show the existence of this monoid by an explicit recursive construction over the size of L. Instead of working with numerical monoids directly, we work in the setting of

If right reversing is complete for the presentation (S, R) and R contains no relation of the form su = sv with s in S and u, v distinct, then the monoid hS | Ri + admits

In a bigon, the strand going to the left at the lower intersection goes to the right at the upper one.. The same happens to the strand connecting the lower and the upper

More examples of monoids of O-type, with the justification and an al- ternative presentation of the associated group; in 1, two-generator symmetric pre- sentations with central

We extend the result obtained in [3] to every Renner monoid: we provide Renner monoids with a monoid presentation and we introduce a length function which extends the Coxeter

On le voit, les variations des récits politiques du passé sont directement liées aux multiples objectifs que ceux-ci cherchent à atteindre : créer du sentiment

We have presented a model of star-forming accretion flows with particular attention to the flow inside the Bondi radius of the central black hole of an elliptical galaxy.. Our

Using Kashiwara’s theory of crystal bases, the author generalized the column presentation for type A and constructed a finite con- vergent presentation of plactic monoid of type C