<<

Canonical bases for the quantum of type Ar and piecewise-linear combinatorics Arkady Berenstein and Andrei Zelevinsky 1 Department of , Northeastern University, Boston, MA 02115

0. Introduction

This work was motivated by the following two problems from the classical representa- tion theory. (Both problems make sense for an arbitrary complex semisimple but since we shall deal only with the Ar case, we formulate them in this generality).

1. Construct a “good” basis in every irreducible finite-dimensional slr+1- Vλ, which “materializes” the Littlewood-Richardson rule. A precise formulation of this problem was given in [3]; we shall explain it in more detail a bit later.

2. Construct a basis in every polynomial representation of GLr+1, such that the maximal element w0 of the Sr+1 (considered as an element of GLr+1)acts on this basis by a permutation (up to a sign), and explicitly compute this permutation. This problem is motivated by recent work by John Stembridge [10] and was brought to our attention by his talk at the Jerusalem Combinatorics Conference, May 1993. We show that the solution to both problems is given by the same basis, obtained by the specialization q = 1 of Lusztig’s canonical basis for the modules over the Drinfeld-Jimbo q-deformation Ur of U(slr+1). More precisely, we work with the basis dual to Lusztig’s, and our main technical tool is the machinery of strings developed in [4]. The solution to Problem 1 appears below as Corollary 6.2, and the solution to Problem 2 is given by Proposition 8.8 and Corollary 8.9. Now let us describe our results and their relationship with the preceding work in more detail. We start with Problem 1. The concept of “good bases” was introduced indepen- dently by K.Baclawski [1] and by I.M.Gelfand-A.Zelevinsky [6]. Technically speaking, in every irreducible slr+1-module Vλ there is a family of subspaces Vλ(β; ν) (their definition and its motivation can be found, e.g., in [3], [4] or in Section 6 below); a basis Bλ in Vλ is good if Bλ ∩ Vλ(β; ν) is a basis in Vλ(β; ν) for every subspace Vλ(β; ν) in this family. As explained in [4], the results by G.Lusztig and M.Kashiwara allow us to construct good bases in the following way. First, each Vλ can be obtained by the specialization q =1from the corresponding irreducible Ur-module, which, with some abuse of notation, we shall denote by the same symbol Vλ. The concept of good bases generalizes to the Ur-modules, and a good basis in the Ur-module Vλ specializes to a good basis in the corresponding slr+1-module. Now consider the algebra A = Ar over the field of rational functions Q(q),

1 Partially supported by the NSF (DMS-9304247).

1 whichisaq-deformation of the ring of regular functions on the maximal unipotent sub- group N+ ⊂ SLr+1.EveryUr-module Vλ has a canonical realization as a subspace of A. There exists a basis B in A (for instance, the dual of Lusztig’s canonical basis) such that Bλ = B ∩ Vλ is a good basis in Vλ for all λ. This basis B and the corresponding bases Bλ are the main objects of study in the present paper.

The basis B is naturally labeled by the so-called Ar-partitions.ByanAr-partition we mean a family of non-negative d =(dij)(i,j)∈Ir , where the index set Ir consists of pairs of integers (i, j) such that 1 ≤ i ≤ j ≤ r. (The set Ir is in a natural bijection with the set of positive roots of type Ar, so we can think of Ar-partitions as partitions of weights into the sum of positive roots.) A labeling of B by Ar-partitions was already used by G.Lusztig (see, e.g., [8], Chapter 42). We use a different approach which allows us to obtain much more explicit results. As an example, let us discuss the above-mentioned “materialization” of the Littlewood-Richardson rule. It is known (see e.g., [3]) that for every three highest weights λ, μ, ν the multiplicity of Vμ in the Vλ ⊗ Vν is equal to the dimension of the subspace Vλ(μ − ν; ν) ⊂ Vλ. So this multiplicity is equal to the number of Ar-partitions d such that the corresponding basis vector bd ∈ B lies in the subspace Vλ(μ − ν; ν) ⊂ Vλ ⊂A. Corollary 6.2 below describes explicitly all such d, thus providing a combinatorial expression for the multiplicity of Vμ in Vλ ⊗ Vν . This expression was earlier obtained in [3], where it was shown to be equivalent to the classical Littlewood-Richardson rule. As for Problem 2 above, we put it in a more general context of studying various symmetries of the bases Bλ. There are several such symmetries, and we compute their action explicitly in terms of Ar-partitions. Our main technical tool is the study of certain continuous piecewise-linear transformations (“transition maps”) acting in the space of Ar- partitions. This is what we mean by “piecewise-linear combinatorics” appearing in the title of this paper. Such combinatorics appeared already in Lusztig’s work; we believe that it constitutes a natural combinatorial framework for the of quantum groups. This framework complements in a nice way the traditional machinery of Young tableaux. Our solution of Problem 2 is a good illustration of the interaction between piecewise- linear and traditional combinatorics. First we compute in terms of Ar-partitions the action on Bλ of a certain involution ηλ : Vλ → Vλ that specializes at q = 1 to the action of w0 (Theorem 7.2 below). Then we translate this description from the language of Ar-partitions to that of Young tableaux (actually, we use an equivalent language of Gelfand-Tsetlin patterns that suits better for our purposes). The final result (Theorem 8.2) is that under a natural parametrization of Bλ by Young tableaux of shape λ, the involution ηλ acts by the well-known Sch¨utzenberger involution. The combinatorial implications of this results are explored by J.Stembridge in [11].

2 The material is organized as follows. For the convenience of the reader we collect in Sections 1 and 2 the necessary material from [4]; all the facts we need about the structure of the Drinfeld-Jimbo algebra Ur and its irreducible modules Vλ are presented in Section 5. In Section 3 we compute the action on the basis B of the natural group of symmetries isomorphic to Z/2Z × Z/2Z. In Section 4 we compute explicitly the so-called exponents of the vectors from B; this calculation is crucial for our derivation of a “piecewise-linear Littlewood-Richardson rule.” The rule itself appears in Section 6. In Section 7 we describe the “twist” of the basis vectors from Bλ under the action of three natural involutive automorphisms of Ur (these automorphisms together with the identity automorphism form another group isomorphic to Z/2Z × Z/2Z). One of these twists is the involution ηλ mentioned above. Finally, in Section 8 we describe the relationship between Ar-partitions and Young tableaux; as a corollary, we show that w0 acts on the canonical basis in every irreducible polynomial GLr+1-module by means of the Sch¨utzenberger involution. The authors are grateful to S.Fomin and J.Stembridge for helpful discussions. This work was partly done during the visit of A. Zelevinsky to the University of Marne La Vallee, France, May - June 1994. He is grateful to J. D´esarm´enien, A. Lascoux, B. Leclerc, J.-Y. Thibon and other members of the Phalanst`ere de Combinatoire Alg´ebrique at Marne La Vallee for their kind hospitality and interest in this work.

1. The algebras A and U+ We fix a positive r and let A denote the with unit over the field of rational functions Q(q) generated by the elements x1,...,xr subject to the relations: xixj = xjxi for |i − j| > 1, (1.1) 2 −1 2 xi xj − (q + q )xixjxi + xjxi =0for|i − j| =1. (1.2) This is a quantum deformation (or q-deformation) of the algebra of polynomial functions on the group of upper unitriangular (r +1)× (r + 1) matrices. We also consider an isomorphic copy U+ of A which is generated by E1,...,Er sat- isfying the same relations as the xi. This is a q-deformation of the universal enveloping algebra of the Lie algebra of nilpotent upper triangular (r +1)× (r + 1) matrices. Both algebras are graded by the semigroup Q+ generated by simple roots α1,...,αr of the of type Ar:wehavedegxi =degEi = αi. The homogeneous components of degree γ will be denoted by A(γ)andU+(γ). These algebras are naturally dual to each other (as graded spaces), according to the following two propositions from [4]. Proposition 1.1. ([4], Proposition 1.1.) There exists a unique action (E,x) → E(x) of the algebra U+ on A satisfying the following properties:

3 (a) (Homogeneity) If E ∈ U+(α),x∈A(γ) then E(x) ∈A(γ − α). (b) (Leibniz formula)

−(γ,αi) Ei(xy)=Ei(x)y + q xEi(y)forx ∈A(γ),y ∈A.

(c) (Normalization) Ei(xj)=δij for i, j =1,...,r. Proposition 1.2. ([4], Proposition 1.2.) (a) If γ ∈ Q+ \{0},andx is a non-zero element of A(γ) then Ei(x) =0 for some i =1,...,r. (b) For every γ ∈ Q+ the mapping (E,x) → E(x) defines a non-degenerate pairing

U+(γ) ×A(γ) →A(0) = Q(q). Here (γ,α) in the Leibniz formula is the usual scalar product on the weight space, so that (αi,αj) is the of type Ar.

2. The canonical basis in A and its string parametrizations We recall from [4] that there is a distinguished basis B in A which is dual to Lusztig’s canonical basis in U+. This basis is a string basis in the terminology of [4]. It is shown in [4] that the elements of B can be labeled by certain integral sequences r(r+1) of length 2 called strings. The string parametrization is associated to every reduced decomposition of w0, the maximal element in the Weyl group. Let us reproduce the definition of strings from [4], Section 2. Let x be a non-zero homogeneous element of A. For each i =1,...,r we set l li(x) = max {l ∈ Z+ : Ei(x) =0 };(2.1)

we call l1(x),l2(x),...,lr(x)theexponents of x. We shall use the following notation:

(top) (li(x)) Ei (x):=Ei (x) (l) (here Ei stands for the divided power, see [4]). Let i =(i1,i2,...,im) be a sequence of indices from {1, 2,...,r} such that no two consecutive indices are equal to each other. We associate to x and i a nonnegative integer vector a(i; x)=(a1,...,am) defined by (top) (top) ··· (top) ak = lik (Eik−1 Eik−2 Ei1 (x)). We call a(i; x)thestring of x in direction i. We abbreviate (top) (top) (top) ··· (top) Ei (x)=Eim Eim−1 Ei1 (x). (top) A − Note that Ei (x) is a non-zero homogeneous element of of degree deg (x) k akαik .

Let W = Sr+1 be the Weyl group of type Ar.Foreachw ∈ W we denote by R(w) the set of all reduced decompositions of w, i.e., the set of sequences i =(i1,i2,...,il) of the minimal possible length l = l(w) such that w is equal to the product of simple ··· reflections si1 si2 sil . We are particularly interested in the reduced decompositions of w0, the maximal element of W .Wedenotem = l(w0)=r(r +1)/2.

4 Proposition 2.1. ([4], Theorems 2.3, 2.4.) For every i =(i1,i2,...,im) ∈ R(w0) we have (top) Ei (b)=1for all b ∈ B. Furthermore, the correspondence b → a(i; b) is a bijection between B and the semigroup CZ(i) of all integral points of some polyhedral convex cone m C(i) ⊂ R+ .

 Proposition 2.2. ([4], Theorem 2.2.) For every i, i ∈ R(w0) there is a piecewise-linear au- m m m  tomorphism i Ti : R → R preserving the Z and such that a(i ; b)=i Ti(a(i; b)) for b ∈ B.

The transition maps i Ti can be computed as follows. It is well-known that any two reduced decompositions of w0 can be transformed into each other by a sequence of elementary transformations of two kinds:

(i1,i,j,i2) → (i1,j,i,i2)for|i − j| > 1, (2.2)

(i1,i,j,i,i2) → (i1,j,i,j,i2)for|i − j| =1, (2.3)

 Proposition 2.3. ([4], Theorem 2.7.) Suppose i = i0, i1,...,ip = i are reduced decom- positions of w0 such that for every t =1,...,p the transition from it−1 to it is of the form (2.2) or (2.3). (a) We have

i Ti = ip Tip−1 ◦···◦i2 Ti1 .

(b) If the transition from it−1 to it is of the form (2.2) with i, j occupying positions k, k+1,

then it Tit−1 leaves all the components of a string a except ak,ak+1 unchanged and changes (ak,ak+1) to (ak+1,ak). (c) If the transition from it−1 to it is of the form (2.3) with i, j, i occupying positions

k, k+1,k+2, then it Tit−1 leaves all the components of a string a except ak,ak+1,ak+2 unchanged and changes (ak,ak+1,ak+2) to

(max (ak+2,ak+1 − ak),ak + ak+2, min (ak,ak+1 − ak+2)).

The most important for us will be the following reduced decomposition of w0:

i(1) = (1; 2, 1; 3, 2, 1; ...; r, r − 1,...,1).

We abbreviate Γ := CZ(i(1)). This semigroup has the following explicit description.

Proposition 2.4. ([4], Theorem 2.5.) The semigroup Γ is the set of all integral sequences

(a11; a22,a12; a33,a23,a13; ...; arr,...,a1r)

5 such that ajj ≥ aj−1,j ≥···≥a1j ≥ 0 for all j =1,...,r.

Note that we have chosen a double indexation for the strings from Γ. Let I = Ir = {(i, j):1≤ i ≤ j ≤ r} be the index set for our numeration. It can be identified with the set of positive roots of type Ar via

(i, j) → αij := αi + αi+1 + ···+ αj. (2.4)

I I I Let Z be the lattice of families (dij)(i,j)∈I of integers indexed by I, and let Z+ ⊂ Z be the semigroup formed by all families (dij) of non-negative integers. We call elements of I I Z+ Ar-partitions (in [7] they were called multisegments). We define a map ∂ :Γ→ Z+ by

∂(a)ij = aij − ai−1,j . (2.5)

For x ∈Awe abbreviate ∂(x):=∂(a(i(1); x)). The following proposition is straightfor- ward. I Proposition 2.5. The map ∂ is a semigroup isomorphism between Γ and Z+. Thus, the → I mapping b ∂(b) is a bijection between B and Z+. Furthermore, if x is a homogeneous element of A and (dij)=∂(x) then the degree of x is equal to (i,j)∈I dijαij.

According to Proposition 2.5, the elements of B can be labeled by Ar-partitions d ∈ I I Z+.Ford ∈ Z+ we shall denote by bd the element of B with ∂(bd)=d. −1 I Clearly, the inverse bijection ∂ : Z+ → Γisgivenby

−1 (∂ (d))ij = d1j + d2j + ···+ dij. (2.6)

3. The action of Z/2Z × Z/2ZonB ∗ ∗ Let x → x be the antiautomorphism of A such that xi = xi for i =1,...,r.Let x → x be the antiautomorphism of A such that xi = xr+1−i for i =1,...,r. Clearly, both maps are involutions and commute with each other. Thus four transformations Id,x→ x∗,x→ x, x → x∗ form a group isomorphic to Z/2Z×Z/2Z. Moreover, it follows from the results of Lusztig (see [8], Section 14.4) that all these transformations preserve I B. So they give an action of Z/2Z × Z/2Z on B. Using the parametrization of B by Z+ I described above, we obtain the action of Z/2Z × Z/2Z on Z+. Three non-trivial elements I ∗ ∗ of Z/2Z×Z/2Z act on Z+ by the mutually commuting involutions d → d ,d→ d, d → d defined by ∗ ∗ ∗ bd = bd , bd = bd, bd = bd∗ . (3.1) We shall describe these involutions quite explicitly. We start with d → d, which turns out to be a permutation of the components dij. We shall use the notation i := r +1− i.

6 ∈ I Proposition 3.1. We have dij = dji for all d Z+. ∈ I † † Proof.Ford =(dij) Z+ we denote temporarily by d the Ar-partition given by dij = dji; so we have to prove that d= d†. We shall use the Poincar´e-Birkhoff-Witt type bases in A constructed in [4]. For each (i, j) ∈ I we set xij =[xi,...[xj−2, [xj−1,xj ]] ...], (3.2)

where [x, y]forx ∈A(γ),y ∈A(γ)istheq-commutator defined by

xy − q(γ,γ )yx [x, y]= (3.3) q − q−1

(see [4], (1.5)); note that in [4] xij was denoted by ti,j+1. The elements xij satisfy the following commutation relations which are special cases of [4], Proposition 3.11:

  [xij,xi,j ]=0fori

[xij,xj+1,k]=xi,k for i ≤ j

Iterating (3.5), we get xij =[...[[xi,xi+1],xi+2] ...,xj]. (3.6)

I For d =(dij) ∈ Z+ we set

d d11 d12 d22 d1r drr x = x11 x12 x22 ···x1r ···xrr . (3.7)

Note that the order of factors here differs from that used in the definition of td in [4], Section 3, which was just the order in Proposition 2.4 above; however, (3.4) implies that the products taken in these two orders differ from each other only by a multiple which is apowerofq. The advantage of the present order is clear from the following result.

d d† I Lemma 3.2. We have x = x for all d ∈ Z+. Proof. First of all, applying the antiautomorphism x → x to both sides of (3.2) and using d d† (3.6) we see that xij = xji. It follows that x is the product of the same factors as x but taken in the different order:

d† d† d† d† d† † d 11 12 1r 22 2r drr x = x11 x12 ···x1r x22 ···x2r ···xrr . (3.8)

The orders in (3.7) and (3.8) differ as follows: the term with xij precedes the term with   xij in (3.7) but goes after it in (3.8) if and only if i

7 by interchanging some commuting terms in the product in (3.8) we can put the factors in the same order as in (3.7). This proves our lemma. I  −1  Consider the following linear order on Z+:wesaythatd ≺ d if and only if ∂ (d ) precedes ∂−1(d) in the lexicographic order. Then the results in [4] imply that the expansion of xd in the basis B has the following form: d n(d) x = q bd + cdd bd , (3.9) d≺d where n(d) is some integer. (To prove (3.9) we notice that ∂(xd)=d, which follows from [4], (3.5) and (3.8), and then apply [4], Proposition 4.1.) Now we can complete the proof of Proposition 3.1. Applying the map x → x to both sides of (3.9) and using Lemma 3.2, we get d† n(d) x = q bd+ cdd bd . (3.10) d≺d

† Comparing (3.10) with the expansion of xd given by (3.9), we conclude that d  d†.In † I † other words, we have d  d for all d ∈ Z+.Butthemapd → d is a weight-preserving I bijection of the set Z+ with itself. Restricting this bijection to every weight component of I † Z+, we obtain a bijection of a finite linearly ordered set with itself such that d  d for all † d. Clearly, such a bijection is the identity map, so we have d† = d for all d. Hence d= d . Proposition 3.1 is proved. ∗ The involution d → d can be expressed in terms of the transition operators i Ti introduced in Section 2. For every sequence i of indices from [1,r]leti∗ denote the sequence obtained from i by replacing each term i by i = r +1−i (leaving the terms in the ∗ ∗ same order). Clearly, if i ∈ R(w0) then i ∈ R(w0); in particular, i(1) ∈ R(w0). Following [4], we denote i(1)∗ by i(r), so

i(r)=(r; r − 1,r; ...;1, 2,...,r). (3.11)

Proposition 3.3. We have

∗ −1 d =(∂ ◦ i(r)Ti(1) ◦ ∂ )(d)

I for all d ∈ Z+. Proof. We start with the following observation. Let E → E∗ and E → E be the antiauto- ∗ morphisms of U+ defined in the same way as the antiautomorphisms x → x and x → x of A ∗ , that is, Ei = Ei, Ei = Ei for i =1,...,r. Then for every two homogeneous elements E ∈ U+,x∈Aofthesamedegreewehave

E(x)=E∗(x∗)=E(x)=E∗(x∗). (3.12)

8 The first equality in (3.12) is proved in [4], Proposition 3.10; the other equalities are proved in exactly the same way. Now since both maps x → x∗ and E → E∗ are automorphisms, the last equality in (3.12) and the definition of strings imply that

a(i∗; x∗)=a(i; x)(3.13) for all i ∈ R(w0) and homogeneous x ∈A. Applying (3.13) to i = i(r)andb = bd for I d ∈ Z+, we obtain our proposition.

Remark. It can be shown that the involution d → d∗ coincides with the multisegment duality ζ, studied recently in [7]. The main result of [7] is an explicit formula for ζ. Comparing this with Propositions 3.1 and 3.3 yields an explicit formula for i(r)Ti(1), which should be helpful for understanding the linearity domains of this piecewise-linear map.

4. The exponents

In this section we compute the exponents (see (2.1) above) of the basis vectors from B.

I Theorem 4.1. For every d =(dij) ∈ Z+ the exponents of the corresponding basis vector bd ∈ B are given by i i−1 lj(bd) = max1≤i≤j ( dhj − dh,j−1)(4.1) h=1 h=1 for j =1,...,r.

Proof. Let a =(a11; a22,a12; ...; arr,...,a1r) be a string from Γ. By slight abuse of notation, we shall write lj(a)forlj(b∂(a)). Taking into account (2.6), we see that (4.1) is equivalent to

lj(a)=max(a1j,a2j − a1,j−1,a3j − a2,j−1,...,ajj − aj−1,j−1). (4.2)

By the definition of strings, lj(a) is the first component of the string iTi(1)(a) for any i ∈ R(w0) which starts with j. So our strategy in proving (4.2) will be to choose some i ∈ R(w0) starting with j, and to compute the first component of iTi(1)(a) by using Proposition 2.3. In doing this, we can assume without loss of generality that j = r (because in computing lj(a) we can just ignore the components aik of a with k>j). ∗ For every sequence i =(i1,...,il) of indices let i =(il,...,i1) denote the same ∗ sequence written in the reverse order. Clearly, if i ∈ R(w0) then i ∈ R(w0) as well.

9 Lemma 4.2. The transformation i(1)∗ Ti(1) can be decomposed into a composition of several transformations of the type described in Proposition 2.3 (b). For every string a =(a11; a22,a12; ...; arr,...,a1r) we have

i(1)∗ Ti(1)(a)=(a11,a22,...,arr; a12,a23,...,ar−1,r; ...; a1,r−1,a2r; a1r). (4.3)

Proof of Lemma 4.2. We shall use the following notation from [4], Section 3: for i

i(1) = (1, 2, 1,...,r, 1), i(1)∗ =(1,r,1,r− 1,...,1, 2, 1).

Let us write i(1) as i(1) = (i(1), r, 1), where i(1) is the element of the same kind as i(1) but with r replaced by r − 1. In order to transform i(1) into i(1)∗, we first ignore the last group r, 1ini(1) and transform i(1) to (i(1)∗, r, 1). Using induction on r,wecan assume that this can be done by a chain of transformations of type (2.2), and that the corresponding transformation of strings is

(i(1)∗,r,1)Ti(1)(a)=(a11,a22,...,ar−1,r−1; a12,a23,...,ar−2,r−1; ...; a1,r−1;

arr,ar−1,r,...,a1r). (4.4)

It remains to transform

(i(1)∗, r, 1) = (1,r− 1, 1,r− 2,...,1, 2, 1, r, 1) into i(1)∗. To do this, we move the term r of the last group r, 1 to the left, interchanging it with its left neighbors until it arrives at the end of the first group 1,r− 1. Then we do the same thing with the terms r − 1,r− 2,...,2 of the last group, moving each of them to the end of the corresponding group 1,r− 2, 1,r− 3,...,1. This sequence of moves transforms the string in (4.4) into that in the right hand side of (4.3), which completes the proof of Lemma 4.2.

Now we can finish the proof of Theorem 4.1. Let us write i(1)∗ as i(1)∗ =(1,r,i(1)∗) and transform it into (1,r,i(1)) by using the transformation i(1)∗ → i(1) inverse to that in Lemma 4.2 (with r replaced by r − 1). Computing the inverse transformation to that given by (4.3) we see that (1,r,i(1))Ti(1)∗ takes the string in (4.3) to the string  (a11,a22,...,arr; a ), where

 a =(a12; a23,a13; ...; ar−1,r,ar−2,r,...,a1r).

10 Now let us recall that our goal is to prove (4.2). Using induction on r we can assume that (4.2) holds for j = r − 1anda replaced by a. This gives

 lr−1(a )=max(a1r,a2r − a1,r−1,a3r − a2,r−1,...,ar−1,r − ar−2,r−1). (4.5)

  Therefore we can find a reduced decomposition of w0 of the form (1,r,i ) such that i begins  with r −1, and the string (1,r,i)T(1,r,i(1))(a11,a22,...,arr; a ) begins with a11,a22,...,arr,   lr−1(a ). To conclude the proof, we concentrate on the first r +1termsof(1,r,i ), which are 1, 2,...,r − 1,r,r − 1. We transform this sequence into r, 1, 2,...,r − 1,r by first performing the operation (r − 1,r,r − 1) → (r, r − 1,r) of type (2.3) and then moving the first term r to the left interchanging it with r − 2,r− 3,...,1. Using Proposition 2.3

(c), (b) we see that these operations transform the string (1,r,i)Ti(1) which begins with   a11,a22,...,arr,lr−1(a ) into the string whose first term is max (lr−1(a ),arr − ar−1,r−1).  By definition, this first term is equal to lr(a), which implies (4.2) since lr−1(a )isgivenby (4.5). This completes the proof of (4.2) and hence of Theorem 4.1. Combining Theorem 4.1 with the results in Section 3, we obtain a formula for the ∗ exponents li(bd) in terms of the Ar-partition d. I ∗ Corollary 4.3. For every d =(dij) ∈ Z+ the exponents of the basis vector bd ∈ B are given by r r ∗ li(bd) = maxi≤j≤r( dik − di+1,k)(4.6) k=j k=j+1 for i =1,...,r. Proof. It follows from (3.13) that ∗ li(x )=li(x)(4.7) ∈A ∗ for i =1,...,r and any homogeneous x . Therefore, we have li(bd)=li(bd)=li(bd). ∗ We see that li(bd) is given by (4.1) with j replaced by i and d replaced by d. Substituting into (4.1) the expressions for the components of d given by Proposition 3.1, we obtain (4.6).

5. The q-analog of slr+1 and its irreducible modules According to Drinfeld and Jimbo, the q-analog of the universal enveloping algebra −1 of slr+1 is the Q(q)-algebra with unit Ur generated by the elements Fi,Ki,Ki ,Ei for i =1,...,r subject to the following relations:

−1 −1 KiKj = KjKi,KiKi = Ki Ki =1; (5.1)

−(αi,αj ) (αi,αj ) KiFj = q FjKi,KiEj = q EjKi for all i, j;(5.2)

11 −1 Ki − Ki EiFj − FjEi = δij ;(5.3) q − q−1

FiFj = FjFi,EiEj = EjEi for |i − j| > 1; (5.4)

2 −1 2 2 −1 2 Fi Fj − (q + q )FiFjFi + FjFi =0,Ei Ej − (q + q )EiEjEi + EjEi =0for|i − j| =1. (5.5) The algebra U+ introduced above, can and will be identified with the Q(q)-subalgebra of Ur generated by all Ei and 1. Let us recall some well-known properties of finite-dimensional Ur-modules. All the proofs can be found in [8]. Let P be the weight lattice of the root system of type Ar, that is,

P = {β ∈ QQ :(β,α) ∈ Z for all α ∈ Q},

where Q is the root lattice and QQ = Q ⊗ Q.LetP+ ⊂ P be the semigroup of dominant weights, that is, the semigroup generated by fundamental weights ω1,...,ωr,wherewe have (ωi,αj)=δij.ForaUr-module V the weight component V (β)ofweightβ ∈ P is defined as

(β,αi) V (β)={v ∈ V : Ki(v)=q v, i =1,...,r}. (5.6)

It is known that every finite-dimensional Ur-module V is diagonalizable, i.e., is the (direct) sum of its weight components. Furthermore, every such module is completely reducible, and the classification of irreducible modules coincides with that for slr+1-modules (see [8], Chapter 6). Thus, to every λ ∈ P+ there corresponds an irreducible finite-dimensional Ur-module Vλ with highest weight λ (so the component Vλ(λ) is one-dimensional and is annihilated by all the Ei). The modules Vλ are non-isomorphic and exhaust all irreducible finite-dimensional Ur-modules. Nowwefixaweightλ = l1ω1 + ···+ lrωr ∈ P+. We shall use an explicit realization of Vλ as a subspace of A given by the following proposition.

Proposition 5.1. (a) The action of U+ on A given by Proposition 1.1 extends to the action of the whole algebra Ur on A which is given by

li−(γ,αi) Ki(x)=q x, (5.7)

l (γ,α )−l q i xxi − q i i xix Fi(x)= (5.8) q − q−1 for x ∈A(γ),i=1,...,r.

12 (li+1) ∗ (b) The elements x ∈Asuch that Ei (x )=0for i =1,...,r form a Ur-submodule of the module A under the action in (a) (here x → x∗ is the antiautomorphism introduced in Section 3). This submodule is isomorphic to Vλ.

This proposition can be deduced from the results in [8] in the following way. In [8], the irreducible module Vλ is realized as a quotient of the Verma module Mλ (see [8], 3.4.5 and Propositions 3.5.6, 6.3.4, 6.3.5). A direct check shows that the Ur-module A described in (a) is obtained from Mλ by passing to the dual module and twisting it by the involutive −1 automorphism ϕ of Ur given by ϕ(Ei)=Fi,ϕ(Fi)=Ei,ϕ(Ki)=Ki (for the definition of the twisting see [8], 3.4.4 or Section 7 below). The same operation of passing to the dual and twisting by ϕ transforms the quotient Vλ of Mλ to the submodule of A described in (b). It remains to observe that this operation transforms Vλ to a module isomorphic to itself. We leave the details of this argument to the reader.

By some abuse of notation, we shall write

(li+1) ∗ Vλ = {x ∈A: Ei (x )=0fori =1,...,r}, (5.9) with the understanding that the action of Ur on Vλ is that in Proposition 5.1 (a). Note that the weight components of Vλ are given by

Vλ(β)=Vλ ∩A(λ − β). (5.10)

In particular, the highest vector of Vλ is just 1 ∈A(0).

6. The canonical basis in Vλ

We retain the notation of the previous sections. So Vλ is an irreducible Ur-module with the highest weight λ = l1ω1 + ···+ lrωr ∈ P+.WesetBλ = B ∩ Vλ,whereB is the canonical basis in A. We recall from Section 2 that the vectors from B are labeled by I Ar-partitions d =(dij) ∈ Z+. Combining (5.9) and (4.6), we get

r r Bλ = {bd : dik − di+1,k ≤ li for 1 ≤ i ≤ j ≤ r}. (6.1) k=j k=j+1

For every β ∈ P and ν = n1ω1 + ···+ nrωr ∈ P+ we set

(ni+1) Vλ(β; ν)={x ∈ Vλ(β):Ei (x)=0fori =1,...,r}, (6.2) where Vλ(β) is the component of weight β in Vλ.

13 Proposition 6.1. For every λ, ν ∈ P+,β∈ P the set Bλ ∩ Vλ(β; ν) is a basis in Vλ(β; ν).

Proof. As in [4], (4.1), for every γ ∈ Q+ and ν ∈ P+ we set

(ni+1) A(γ; ν)={x ∈A(γ):Ei (x)=0fori =1,...,r}.

By [4], Proposition 4.6, each subspace A(γ; ν) ⊂Ais spanned by its intersection with B. Since B is invariant under the antiautomorphism x → x∗, every subspace A(γ; ν)∗ is also spanned by its intersection with B. This implies our statement since, in view of (5.9) and (5.10), we have ∗ Vλ(β; ν)=A(λ − β; ν) ∩A(λ − β; λ) .

In view of Proposition 2.5 and (4.1), we can reformulate Proposition 6.1 as follows.

Corollary 6.2. The space Vλ(β; ν) has as a basis the set of elements bd,whered runs over all Ar-partitions satisfying three conditions: dijαij = λ − β, (6.3) i,j

r r dik − di+1,k ≤ li for 1 ≤ i ≤ j ≤ r. (6.4) k=j k=j+1 i i−1 dhj − dh,j−1 ≤ nj for 1 ≤ i ≤ j ≤ r. (6.5) h=1 h=1

Corollary 6.2 allows us to “materialize” the Littlewood- Richardson rule along the lines of [3]. To do this we notice that the specialization q = 1 makes Vλ into an irreducible slr+1-module with highest weight λ, which we will (with some abuse of notation) denote also by Vλ. More precisely, it is known (see [8], Chapter 22) that all the matrix entries of the operators Ei and Fi (acting on Vλ) in the basis Bλ are rational functions in q regular at q = 1. Therefore, we can define the operators ei and fi by specializing the matrices of Ei and Fi at q = 1, and let hi act on each weight component Vλ(β) by multiplication by (β,αi). Then the relations (5.1)-(5.5) imply that the ei,fi and hi satisfy the commutation relations among the Cartan generators of slr+1. This makes the C-space with the basis Bλ an slr+1-module. Under this specialization, the subspace Vλ(β; ν) ⊂ Vλ (orratheritsC-form) becomes ni+1 the space of vectors of weight β in Vλ annihilated by the operators ei ,i=1,...,r.Itis well-known that the dimension of this space is equal to the multiplicity of the irreducible slr+1-module Vν+β in the tensor product Vλ ⊗ Vν (see [3]). Thus, Corollary 6.2 implies that this multiplicity is equal to the number of Ar-partitions d satisfying (6.3)-(6.5). This

14 statement was established in [3] in a combinatorial way, essentially by showing that it is equivalent to the classical Littlewood-Richardson rule. Proposition 6.1 and Corollary 6.2 provide us with a representation-theoretic proof of this result. ∗ Note also that the map x → x induces an isomorphism of vector spaces Vλ(λ − γ; ν) ∗ and Vν (ν − γ; λ) for every λ, ν ∈ P+,γ∈ P . Thus, the involution d → d on Ar-partitions provides a bijective proof of the fact that our expressions for the multiplicity of Vλ+ν−γ in Vλ ⊗ Vν and the multiplicity of Vλ+ν−γ in Vν ⊗ Vλ give the same answer.

7. Twisting Bλ by the automorphisms of Ur The commutation relations (5.1)-(5.5) imply that there exist three Q(q)-linear invo- lutive automorphisms ϕ, ψ and η of the algebra Ur acting on the generators as follows: −1 ϕ(Ei)=Fi,ϕ(Fi)=Ei,ϕ(Ki)=Ki ;(7.1)

ψ(Ei)=Ei,ψ(Fi)=Fi,ψ(Ki)=Ki;(7.2) −1 η(Ei)=F ,η(Fi)=E ,η(Ki)=K , (7.3) i i i where i = r +1− i. Clearly, these three automorphisms together with the identity au- tomorphism form a group isomorphic to Z/2Z × Z/2Z. It turns out that each of these automorphisms induces some transformation (“twist”) of the bases Bλ; in this section we shall compute this twist explicitly. The general setup is as follows. If V is a module over an associative algebra U and σ is an automorphism of U, then the twisted U-module σV is the same vector space V but with the new action u ∗ v = σ−1(u)v, u ∈ U, v ∈ V . Clearly, στ V = σ(τ V ) for every two automorphisms σ, τ of U. Furthermore, if V is a simple U-module then so is σV .In σ particular, if U = Ur and V = Vλ then Vλ is isomorphic to Vσ(λ) for some highest weight σ(λ). Thus there exists an isomorphism of vector spaces σλ : Vλ → Vσ(λ) such that

σλ(uv)=σ(u)σλ(v),u∈ Ur,v∈ Vλ.

Clearly, σλ is unique up to a scalar multiple. It follows that the operator στ(λ)τλ is proportional to (στ)λ for every two automorphisms σ and τ of Ur. Returning to our situation, it follows at once from (7.1) - (7.3) that

ϕ(λ)=ψ(λ)=−w0(λ),η(λ)=λ, (7.4)

where w0 is the maximal element of the Weyl group. We recall that every module Vλ is canonically realized as a subspace in A, so that the highest vector in Vλ is 1. We denote by low bλ the lowest weight vector in Vλ, normalized by the condition that it lies in Bλ = B ∩Vλ. Now we normalize each of the maps ϕλ,ψλ and ηλ by the requirement that low low ϕλ(1) = b−w0(λ),ψλ(1) = 1,ηλ(1) = bλ (7.5)

(of course, we also set Idλ to be the identity map of Vλ).

15 Proposition 7.1.

(a) Each of the maps ϕλ and ψλ sends Bλ to B−w0(λ), while ηλ sends Bλ to Bλ. (b) For every two (not necessarily distinct) elements σ, τ of the group {Id,ϕ,ψ,η} we have (στ)λ = στ(λ)τλ. Part (a) of the proposition is proved in [8], Proposition 21.1.2. Part (b) follows from (a) and the fact that στ(λ)τλ and (στ)λ are always proportional to each other. ∈{ } Using Proposition 7.1 (a), we shall write σλ(bd)=bσλ(d) for all σ Id,ϕ,ψ,η and all bd ∈ Bλ. So the map d → σλ(d)onAr-partitions is well-defined for d such that bd ∈ Bλ, that is, for d satisfying (6.4).

Theorem 7.2. The mappings ϕλ,ψλ and ηλ act on Ar-partitions as follows: r r ϕλ(d)j+1−i,j = li − ( dik − di+1,k); (7.6) k=j k=j+1

∗ ψλ(d)r+1−j,r+1−i = dij;(7.7) i i−1 ∗ ∗ ηλ(d)j+1−i,r+1−i = lj − ( dhj − dh,j−1)(7.8) h=1 h=1 for all i, j such that 1 ≤ i ≤ j ≤ r. Proof. First of all, let us show that (7.8) follows from (7.6) and (7.7). Indeed, in view of

Proposition 7.1 (b) and (7.4), we have ηλ(d)=ϕ−w0(λ)ψλ(d), so (7.6) and (7.7) imply

ηλ(d)j+1−i,r+1−i = ϕ−w0(λ)ψλ(d)i+1−j,i r r − − = lj ( ψλ(d)j,k ψλ(d)j+1,k) k=i k=i+1 r r ∗ ∗ j − − = l ( dk,j dk,j−1). k=i k=i+1

Substituting h = k in the last summation yields (7.8). → To prove (7.7), it is enough to show that the map ψλ : Vλ V−w0(λ) is just the ∗ restriction to Vλ of the automorphism x → x of the algebra A (see Proposition 3.1 above). Remembering the definitions and the fact that x → x∗ preserves B, it is easy to see that low low ∗ ∗ ∈ + ∈A this map sends bλ to b−w0(λ) and has the property ux = ψ(u)x for u U ,x .It ∗ follows that ψλ(x)=x , which implies (7.7). It remains to prove (7.6). Fix b = bd ∈ Bλ ∩ Vλ(β), and consider the strings

a =(a11; a22,a12; ...; arr,...,a1r)=a(i(1); b),

16 − − − − − − a =(a11; a22,a12; ...; arr,...,a1r)=a(i(1); ϕλ(b)). By the definition (2.5), we have

− − dij = aij − ai−1,j ,ϕλ(d)j+1−i,j = aj+1−i,j − aj−i,j, (7.9)

so it will be enough for us to express the string a− through a. For j =1,...,r we consider Uj as the subalgebra of Ur generated by the elements ±1 Ei,Fi and Ki for i =1,...,j.Wedenotebyi(1,j) the initial subword (1; 2, 1; ...; j, j − 1,...,1) of i(1) (so i(1) itself is now denoted i(1,r)). Clearly, i(1,j) ∈ R(w0(j)) is a reduced decomposition of the element w0(j) ∈ W which can be identified with the maximal element of the Weyl group of Uj. We set (a1j ) (a2j ) (ajj) (a12) (a22) (a11) b(j)=E1 E2 ···Ej ···E1 E2 E1 (b);

(top) in the notation of Section 2, we have b(j)=Ei(1,j)(b). We also set

(a− ) (a− ) (a− ) − − − − 1j 2j jj (a12) (a22) (a11) b (j)=F1 F2 ···Fj ···F1 F2 F1 (b). (7.10)

In view of (7.1), we have − −1 (top) b (j)=ϕλ (Ei(1,j)(ϕλ(b))). (7.11)

− (top) It follows that we can also write b (j)=Fi(1,j)(b), with the obvious meaning that each (ai) power of the type Fi appearing in (7.10) is the maximal possible power of Fi which still produces a non-zero vector. We shall show that for each j the vectors b(j)andb−(j) are, respectively, the highest and lowest weight vectors of the same irreducible Uj-submodule in Vλ. We need the following results from [4] which we reproduce here for the convenience of the reader. Lemma 7.3. ([4], Theorem 2.2, Proposition 6.1). Let b ∈ B,w ∈ W and i ∈ R(w). Then (top) the element Ei (b) belongs to B, depends only on w (not on the choice of a reduced decomposition of w), and is annihilated by all Ei such that l(wsi)

Lemma 7.3 implies that Eib(j)=0fori =1,...,j,sob(j) is a highest weight vector of − some irreducible Uj-submodule of Vλ. Using (7.11), we see also that b (j) is a lowest weight − vector of some irreducible Uj-submodule of Vλ. To show that b(j)andb (j) generate the (top) − same Uj-submodule, it is enough to see that b(j)=Ei(1,j)(b (j)). In other words, we have to check the equality (top) (top) (top) Ei(1,j) ◦ Fi(1,j) = Ei(1,j), (7.12)

where both sides are considered as operators acting on Bλ. To prove (7.12) we first notice (top) (top) (top) that Ei ◦ Fi = Ei for i =1,...,r; this follows from the commutation relations

17 (5.2), (5.3) between Ei,Fi and Ki by a standard argument from the representation theory of sl2.Nextweprovethat (top) (top) (top) Ei(1,j) ◦ Fi = Ei(1,j) (7.13) for i =1,...,j. To see this, we notice that there exists a reduced decomposition of w0(j)  (top) (top) starting with i, i.e., having the form (i, i ); in view of Lemma 7.3, we have Ei(1,j) = E(i,i) = (top) (top) Ei ◦ Ei ,so

(top) (top) (top) (top) (top) (top) (top) (top) Ei(1,j) ◦ Fi = Ei ◦ Ei ◦ Fi = Ei ◦ Ei = Ei(1,j),

(top) proving (7.13). Finally, (7.13) obvioulsy implies (7.12) since Ei(1,j) “swallows” all the (top) (top) factors Fi occuring in Fi(1,j). Since b(j)andb−(j) are the highest and lowest weight vectors of the same irreducible Uj-module, their weights with respect to Uj are obtained from each other by the action of w0(j). To write down this statement explicitly, we need some notation. Let P (j) be the weight lattice for Uj (so the lattice P is P (r)). To distinguish the fundamental    weights and simple roots of Uj from those of Ur, we shall denote them by ω1,ω2,...,ωj    and α1,α2,...,αj . Clearly, the natural projection pj : P → P (j)actsasfollows:  ωi if 1 ≤ i ≤ j pj(ωi)= . (7.14) 0ifj

We also have

    w0(j)(ωi)=−ωj+1−i,w0(j)(αi)=−αj+1−i for i =1,...,j. (7.15)

Clearly, the Uj-weight of b(j) is equal to

j j  pj(β)+ ( aik) · αi; i=1 k=i

− similarly, the Uj-weight of b (j) is equal to

j j −  pj(β) − ( aik) · αi. i=1 k=i

Therefore, we have the equality

j j j j  −  pj(β)+ ( aik) · αi = w0(j)[pj(β) − ( aik) · αi] i=1 k=i i=1 k=i

18 j j −  = w0(j)pj(β)+ ( aj+1−i,k) · αi. (7.16) i=1 k=j+1−i  Taking the scalar product of both sides of (7.16) with ωi for i =1,...,j, we obtain

j j − aj+1−i,k = aik + cij(β), (7.17) k=j+1−i k=i where   cij(β)=(pj(β),ωi) − (w0(j)pj(β),ωi). (7.18)

The rest of the proof is a formal calculation deducing (7.6) from (7.17), (7.18). First, we rewrite (7.18) in a more convenient form. Using (7.15) and the fact that w0(j) preserves the scalar product in P (j), we obtain

  cij(β)=(pj(β),ωi + ωj+1−i).

An easy check shows that

i      ωi + ωj+1−i = (αh + αh+1 + ···+ αh+j−i). h=1

Using (7.14), we can rewrite (7.18) as

i cij(β)=(β, (αh + αh+1 + ···+ αh+j−i)). (7.19) h=1

Now we subtract from (7.17) the similar equality obtained from it by replacing (i, j) with (i − 1,j− 1). Using (7.19), we can rewrite the resulting equality as follows:

j j−1 − aj+1−i,j = aik − ai−1,k + cij(β) − ci−1,j−1(β) k=i k=i−1

j−1 j = aij − ai−1,i−1 + dik +(β, αk). (7.20) k=i k=i Subtracting from (7.20) the similar inequality obtained from it by replacing (i, j) with (i +1,j), and taking into account (7.9), we obtain

j−1 j ϕλ(d)j+1−i,j = aii − ai−1,i−1 + dik − di+1,k +(β,αi). (7.21) k=i k=i+1

19 The final step is to compute (β,αi). In view of (5.10), we have

r r β = λ − ( aik) · αi, i=1 k=i

which implies r r r (β,αi)=li + ai−1,k − 2 aik + ai+1,k k=i−1 k=i k=i+1 r r = li + ai−1,i−1 − aii − dik + di+1,k. k=i k=i+1

Substituting this expression into (7.21) and performing the obvious cancellation we obtain (7.6). Theorem 7.2 is proved.

8. The Sch¨utzenberger involution

In this section we give a combinatorial description of the involution ηλ in terms of Young tableaux and Gelfand-Tsetlin patterns. Here λ = l1ω1 + ···+ lrωr is a fixed highest weight for slr+1. We associate with λ a partition Λ = (λ1 ≥ λ2 ≥ ··· ≥ λr+1 ≥ 0) of length ≤ r +1,where λr+1 is an arbitrary non-negative integer, and li = λi − λi+1 for i =1,...,r. Let us recall some well-known combinatorial definitions. We identify Λ with its diagram (denoted by the same letter)

Λ={(i, j) ∈ Z × Z :1≤ i ≤ r +1, 1 ≤ j ≤ λi}.

By an Ar-tableau of shape Λ we shall mean a map τ :Λ→ [1,r+1] satisfying the conditions

τ(i, j +1)≥ τ(i, j),τ(i +1,j) >τ(i, j)(8.1) for all (i, j) ∈ Λ; here and in the sequel we adopt the convention that τ(i, j)=+∞ for i>r+1,j ≥ 1or1≤ i ≤ r +1,j>λi. (In the combinatorial literature the tableaux satisfying (8.1) are often called semistandard.) The weight of an Ar-tableau τ is an integral vector β =(β1,...,βr+1) defined by −1 r βi =#τ (i). We denote by the same letter β the slr+1-weight i=1(βi − βi+1)ωi. The language of Ar-tableaux is equivalent to that of Gelfand-Tsetlin patterns, which will be more convenient for us. By a GT-pattern of highest weight Λ we mean an array of integers π =(πij)1≤i≤j≤r+1 such that πi,r+1 = λi for i =1,...,r+1and

πi,j+1 ≥ πij ≥ πi+1,j+1 (8.2)

20 for 1 ≤ i ≤ j ≤ r. Such a pattern is displayed as a triangular array ⎛ ⎞ π14 π24 π34 π44 ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ π13 π23 π33 ⎟ ⎜ ⎟ π = ⎜ ⎟ . ⎜ ⎟ ⎝ π12 π22 ⎠

π11

Clearly, the numbers in each row are weakly decreasing: π1j ≥ π2j ≥···≥πjj.Theweight of a GT-pattern π is an integral vector β =(β1,...,βr+1) defined by

β1 + β2 + ···+ βj = π1j + π2j + ···+ πjj for j =1,...,r+1. Let YTΛ denote the set of all Ar-tableaux of shape Λ, and GTΛ denote the set of all GT-patterns of highest weight Λ. It is well-known that the sets YTΛ and GTΛ can be identified with each other by the following weight-preserving bijection τ → π(τ)=(πij):

πij =#{s :1≤ s ≤ λi,τ(i, s) ≤ j};(8.3) indeed, it is easy to see that (8.3) transforms the conditions (8.1) to (8.2). Now we introduce the Sch¨utzenberger involution η : GTΛ → GTΛ. It was first defined in [9] by means of a beautiful combinatorial sometimes called the evacuation.We shall use an equivalent definition due to E.Gansner [5]. It is given in terms of the so-called Bender-Knuth involutive operators t1,t2,...,tr acting on Ar-tableaux. Translating them into the language of GT-patterns, we arrive at the following definition. For j =1,...,r and π =(πij) ∈ GTΛ we define the pattern tj(π) ∈ GTΛ by tj(π)ik = πik for k = j, tj(π)ij = min(πi,j+1,πi−1,j−1)+max(πi+1,j+1,πi,j−1)−πij. (8.4)

(In view of (8.2), if we fix all the components πik with k = j then each πij takes the values in the segment [max(πi+1,j+1,πi,j−1), min(πi,j+1,πi−1,j−1)]; the transformation (8.4) is just the reflection of πij in the midpoint of this segment.) Clearly, each tj is an involutive map GTΛ → GTΛ. As shown in [5], the Sch¨utzenberger involution η : GTΛ → GTΛ canbedefinedby

η := (t1 ···tr)(t1 ···tr−1) ···(t1t2)t1;(8.5) the fact that η is an involution follows readily from the obvious relations

2 tj =1,titj = tjti for |i − j| > 1.

21 (The group generated by the piecewise-linear automorphisms t1,...,tr was studied in [2].) I We define the mapping ∂ : GTΛ → Z+ by the formula

∂(π)ij = πi,j+1 − πij (1 ≤ i ≤ j ≤ r). (8.6)

Comparing (8.6) and (8.3) we see that

∂(π(τ))ij =#{s :1≤ s ≤ λi,τ(i, s)=j +1} (8.7)

(1) for any τ ∈ YTΛ; in the notation of [7], we have ∂(π(τ)) = d (τ). The following proposition is an easy consequence of (8.6) and (8.7) combined with (6.1) and (6.3) (cf. [7], (4.4)).

I Proposition 8.1. The map ∂ is a bijection of GTΛ with the set {d ∈ Z+ : bd ∈ Bλ}.In other words, the canonical basis Bλ in Vλ can be parametrized by GT-patterns of highest weight Λ via π → b∂(π). Furthermore, if π ∈ GTΛ has weight β then b∂(π) ∈ Vλ(β), i.e., the parametrization π → b∂(π) is weight-preserving.

Now we can state the main result of this section.

Theorem 8.2. For any π ∈ GTΛ we have

ηλ(b∂(π))=b∂(η(π)). (8.8)

In other words, under the parametrization of basis vectors by GT-patterns, the invo- lution ηλ acts on patterns as the Sch¨utzenberger involution. In order to deduce Theorem 8.2 from Theorem 7.2, we shall translate the operations I I tj into the language of Ar-partitions. We define the maps R1,R2,...,Rr : Z+ → Z+ in I the following way. For d =(dij) ∈ Z+ and j =1,...,r we define Rj(d)by

Rj(d)ik = dik for k = j, j − 1;

Rj(d)ij = min (dij,di−1,j−1)+[di,j−1 − di+1,j ]+;(8.9)

Rj(d)i,j−1 = min (dij,di−1,j−1)+[di+1,j − di,j−1]+, where [x]+ =max(0,x). Here we use the convention that d0,j−1 that can appear under the minimum sign in (8.9) is +∞, and [dj,j−1 − dj+1,j ]+ = 0. In particular, R1 is just the identity map. −1 Note that each Rj is invertible, and the inverse map Rj is given by similar formulas

−1 Rj (d)ik = dik for k = j, j − 1;

22 −1 Rj (d)ij = min (dij,di,j−1)+[di−1,j−1 − di−1,j ]+;(8.10) −1 Rj (d)i−1,j−1 = min (dij,di,j−1)+[di−1,j − di−1,j−1]+

(with the conventions similar to the above). The check that the maps defined by (8.10) and (8.9) are indeed inverse to each other, is straightforward (it uses the obvious identity [x]+ − [−x]+ = x). The most important for us will be the following composition of the maps Rj, which is reminiscent of the definition (8.5) of the Sch¨utzenberger involution η:

ρ := (R1 ···Rr)(R1 ···Rr−1) ···(R1R2)R1. (8.11)

We shall express both η and the involution d → d∗ in terms of the map ρ. For this  I we need another map ∂ : GTΛ → Z+, whose definition is similar to (8.6):

 ∂ (π)ij = πij − πi+1,j+1 (1 ≤ i ≤ j ≤ r). (8.12)

Comparing (8.12) and (8.3) we see that

 ∂ (π(τ))ij =#{s :1≤ s ≤ λi,τ(i, s) ≤ j, τ(i +1,s) ≥ j +2} (8.13)

 (2) for any τ ∈ YTΛ; in the notation of [7], we have ∂ (π(τ)) = d (τ). (1) A direct calculation shows that for every π ∈ GTΛ the Ar-partitions d = ∂(π)and d(2) = ∂(π) are related as follows (cf. [7], (4.5), (4.6)): r r (2) (1) (1) dij = li − ( di,k − di+1,k); (8.14) k=j k=j+1

i i−1 (1) (2) (2) dj+1−i,r+1−i = lj − ( dj+1−h,r+1−h − dj−h,r+1−h). (8.15) h=1 h=1 Theorem 8.3.  (a) The composition of ∂ with the Sch¨utzenberger involution η : GTΛ → GTΛ is equal to

∂η = ρ∂. (8.16)

I ∗ (b) For every d ∈ Z+ the Ar-partitions ρ(d) and d are related as follows:

∗ ρ(d)j+1−i,r+1−i = dij. (8.17)

Remark. We have already mentioned that d → d∗ coincides with the involution ζ from [7]. In fact, Theorem 8.3 is equivalent to Theorem 4.4 from [7], with ζ replaced by d → d∗.

23 Before proving Theorem 8.3, let us show that together with Theorem 7.2 it implies  Theorem 8.2. Let π ∈ GTΛ,andd = ∂(π). By (8.16), ∂ (η(π)) = ρ(d); so (8.17) gives

 ∗ ∂ (η(π))j+1−i,r+1−i = dij.

Substituting this expression into the right hand side of (8.15), we obtain

i i−1 ∗ ∗ ∂(η(π))j+1−i,r+1−i = lj − ( dhj − dh,j−1). h=1 h=1

Comparing this with (7.8), we see that ηλ(bd)=b∂(η(π)), which proves Theorem 8.2.

 ε I Proof of Theorem 8.3 (a). We include ∂ and ∂ into a family of mappings ∂ : GTΛ → Z+, where ε is a sign vector ε =(ε1,...,εr), each εj being either + or −.For1≤ i ≤ j ≤ r we set  ε ∂ (π)ij if εj =+. ∂ (π)ij = (8.18) ∂(π)ij if εj = −.

In particular, we have ∂(−,−,...,−) = ∂, ∂(+,+,...,+) = ∂. If a sign vector ε is such that εj = −,εj−1 = + then we denote by sj(ε) the sign vector obtained from ε by switching εj to+andεj−1 to − (in particular, s1(ε)makes sense if ε1 = − and is then obtained from ε by switching this − to +).

Lemma 8.4. If a sign vector ε is such that sj(ε) makes sense then we have

sj (ε) ε ∂ tj = Rj∂ . (8.19)

ε Proof. Let π =(πij) ∈ GTΛ,andd = ∂ (π), so that

dij = πi,j+1 − πij,di,j−1 = πi,j−1 − πi+1,j . (8.20)

It follows that πi,j−1 − πi+1,j+1 = di,j−1 − di+1,j . (8.21)

Using (8.4), (8.20), (8.21) and (8.9), we get

sj (ε) ∂ tj(π)ij = tj(π)ij − tj(π)i+1,j+1

= min(πi,j+1,πi−1,j−1) + max(πi+1,j+1,πi,j−1) − πij − πi+1,j+1

= min(πi,j+1 − πij,πi−1,j−1 − πij)+[πi,j−1 − πi+1,j+1]+

= min(dij,di−1,j−1)+[di,j−1 − di+1,j ]+ = Rj(d)ij.

24 s (ε) The equality ∂ j tj(π)i,j−1 = Rj(d)i,j−1 is proved in exactly the same way. Finally, for k = j, j − 1wehave

sj (ε) sj (ε) ∂ tj(π)ik = ∂ (π)ik = dik = Rj(d)ik.

This completes the proof of the lemma. To complete the proof of Theorem 8.3 (a), we notice that the sign vector (+, +,...,+) canbeobtainedfrom(−, −,...,−) by the following chain of transformations:

(+, +,...,+) = (s1 ···sr)(s1 ···sr−1) ···(s1s2)s1(−, −,...,−). (8.22)

Using this chain of transformations and applying Lemma 8.4 in each step, we obtain

(−,−,...,−) (R1 ···Rr)(R1 ···Rr−1) ···(R1R2)R1∂

(+,+,...,+) = ∂ (t1 ···tr)(t1 ···tr−1) ···(t1t2)t1,

which proves (8.16). Proof of Theorem 8.3 (b). In view of Proposition 3.3, the involution d → d∗ is closely related to the transition operation i(r)Ti(1) on strings. We start with studying the transition from i(1) to i(r) in more detail. To do this, we include i(1) and i(r) into a family of reduced decompositions i(ε) ∈ R(w0), labeled by the same sign vectors ε as in part (a) above. We (1) (2) (r) (j) shall write a reduced decomposition i ∈ R(w0)asi =(i ; i ; ...; i ), where each i is a sequence of indices of length j. We shall also use the notation i, j introduced in the proof of Lemma 4.2 above. In this notation, if ε =(ε1,...,εr) is a sign vector then i(ε) ∈ R(w0) is uniquely determined by the following properties: (j) (1) Each i(ε) has the form k, k + j − 1 for some k if εj =+,andk, k − j +1forsome (r) (r) k if εj = − (in particular, i(ε) = 1,r if εr =+,andi(ε) = r, 1ifεr = −). (j) (2) Let j ≥ 2. If i(ε) = k, k + j − 1 (i.e., εj = +) then we have

(j−1) εj−1 =+ ⇒ i(ε) = k +1,k+ j − 1,

(j−1) εj−1 = −⇒i(ε) = k + j − 1,k+1.

(j) If i(ε) = k, k − j + 1 (i.e., εj = −) then we have

(j−1) εj−1 =+ ⇒ i(ε) = k − j +1,k− 1,

(j−1) εj−1 = −⇒i(ε) = k − 1,k− j +1.

25 In particular, we have i(1) = i(−, −,...,−), i(r)=i(+, +,...,+). It is easy to check that all i(ε) are indeed reduced decompositions of w0 (for instance, using induction on r). Clearly, the map ε → i(ε) is two-to-one: if ε and ε differ only in the first component then i(ε)=i(ε). (1) (2) (r) (j−1) (j) Suppose i =(i ; i ; ...; i ) ∈ R(w0) is such that for some j the pair (i ; i )has the form (k − j +1,k− 1; k, k − j + 1). We denote by sj(i) the sequence obtained from i (j−1) (j) by replacing (i ; i )with(k, k − j +2;k − j +1,k). It is clear that sj(i) ∈ R(w0) (this follows from the fact that both (k − j +1,k− 1; k, k − j + 1) and (k, k − j +2;k − j +1,k) are reduced decompositions of the same element in Sr+1). The operation i → sj(i)iscon- sistent with the operation ε → sj(ε) on sign vectors introduced above: the definitions imply at once that sj(i(ε)) = i(sj(ε)), (8.23)

whenever sj(ε) makes sense (we use the convention that s1 is the identity operation on reduced decompositions). Comparing (8.22) and (8.23) we see that

i(r)=(s1 ···sr)(s1 ···sr−1) ···(s1s2)s1(i(1)). (8.24)

Our next step is to lift the chain of transformations (8.24) to the level of strings. We shall write vectors from Rm as x =(x(1); ...; x(r)), where each x(j) belongs to Rj.

Proposition 8.5. If i ∈ R(w0) is such that sj(i) makes sense then the transition map m → m sj (i)Ti : R R depends only on j, not on the choice of i. This map leaves unchanged all the components x(k) of x ∈ Rm except x(j) and x(j−1); furthermore, the (j − 1)-st and (j) (j−1) j-th components of sj (i)Ti(x) depend only on x and x . Proof. The proposition follows at once from the description of transition maps given (1) (r) by Proposition 2.3. Indeed, sj(i) is obtained from i =(i ; ...; i ) by transforming (i(j−1); i(j))=(k − j +1,k− 1; k, k − j +1)to(k, k − j +2;k − j +1,k). In order to − − − → compute sj (i)Ti we have to decompose the transformation (k j +1,k 1; k, k j +1) (k, k − j +2;k − j +1,k) into a sequence of transformations of type (2.2) and (2.3), and then compose the corresponding mappings in Proposition 2.3 (b), (c). Clearly, this com- position does not depend on k and has all the properties claimed in Proposition 8.5. m → m In view of Proposition 8.5, we shall denote the map sj (i)Ti : R R simply by Tj; in particular, T1 is the identity map. Combining Proposition 2.3 (a) and (8.24), we obtain

i(r)Ti(1) =(T1 ···Tr)(T1 ···Tr−1) ···(T1T2)T1. (8.25)

The last step in the proof of Theorem 8.3 (b) is to relate the maps Tj on strings m with the maps Rj on Ar-partitions. We recall that the semigroup Γ ⊂ R consists of (1) (r) (j) j x =(x ; ...; x ) such that for j =1,...,r we have x ∈ Z+ and the components of

26 (j) (j) (j) (j) x are weakly decreasing: x1 ≥ x2 ≥ ··· ≥ xj ≥ 0 (see Proposition 2.4). For every ε I sign vector ε we define the map ∂ :Γ→ Z+ by (j) (j) ε xi − xi+1 if εj =+; ∂ (x)ij = (j) (j) (8.26) xj+1−i − xj+2−i if εj = −,

(j) ε with the convention xj+1 = 0. Clearly, all ∂ are semigroup isomorphisms. In particular, ∂(−,−,...,−) is the map ∂ in (2.5). The following lemma is completely analogous to Lemma 8.4.

Lemma 8.6. If a sign vector ε is such that sj(ε) makes sense then we have

sj (ε) ε ∂ Tj = Rj∂ (8.27)

I (as mappings Γ → Z+). Proof. An easy calculation using the strategy described in the proof of Proposition 8.5, (j) (j−1) shows that Tj acts on the components x and x in the following way:

(j) (j−1) (j−1) (j) (j) Tj(x)i = min (xi−1 ,xi + xj+1−i − xj+2−i),

(j−1) (j) (j) (j−1) (j−1) Tj(x)i =max(xi+1,xi + xj+1−i − xj−i ). (8.28) (In fact, a part of this calculation was already done in the end of the proof of Theorem 4.1.) Now (8.27) follows from (8.28), (8.26) and (8.9) by a straightforward calculation com- pletely analogous to that in the proof of Lemma 8.4. We leave the details to the reader. Now we can complete the proof of Theorem 8.3 (b). As in the proof of part (a), using repeatedly Lemma 8.6, we deduce from (8.25) and (8.22) that

(+,+,...,+) ∂ i(r)Ti(1) = ρ∂. (8.29)

I ∗ Comparing (8.29) with Proposition 3.3, we see that for every d ∈ Z+ the Ar-partitions d and ρ(d) are related by ∗ d ij = ρ(d)j+1−i,j .

It follows that ∗ ∗ dij = d r+1−j,r+1−i = ρ(d)j+1−i,r+1−i, which is the identity (8.17). Theorems 8.3 and 8.2 are proved.

Remark 8.7. The above calculations allow us to relate our transition maps i Ti with i m m Lusztig’s piecewise-linear automorphisms Ri : R → R (the definition of Lusztig’s maps

27 can be found, e.g., in [8], 42.1.3, 42.2.1). Namely, let us define the linear automorphism δ : Rm → Rm by the formula

(j) (j) (j) δ(x)i = xj+1−i − xj+2−i

I (so δ(x) is obtained from the vector d = ∂(x) ∈ R by arranging its components dij in the order (d11,d12,d22,...,d1r,...,drr); cf. (8.26)). Then an easy calculation using Lemma 8.6 shows that for any two sign vectors ε, ε we have

i(ε) −1 Ri(ε) = δ ◦ i(ε)Ti(ε) ◦ δ . (8.30)

This formula shows in fact that the reduced decompositions i(ε) are rather special. It is almost obvious from the definitions that there exists a unique family of piecewise-linear m m automorphisms δi : R → R (i ∈ R(w0)) such that δi(1) = δ and

i −1 Ri = δi ◦ i Ti ◦ δi

 for all i, i ∈ R(w0). In this notation, (8.30) means that δi(ε) = δ for every sign vector ε.In general, the maps δi are not even linear; the simplest example is given by i =(1, 3, 2, 1, 3, 2) for r = 3. Finding an explicit formula for δi is an interesting problem in piecewise-linear combinatorics. Our last result is an application of Theorem 8.2 to the classical representation theory obtained by the specialization q = 1. Having in mind potential combinatorial aplications, we prefer to speak about polynomial representations of the group GLr+1.LetVΛ be the polynomial representation of GLr+1 corresponding to a partition Λ. As an slr+1-module, VΛ is just an irreducible module Vλ. Therefore, as explained in the end of Section 6, we can view Bλ as a basis in VΛ. On the other hand, an element w0 ∈ Sr+1 can be viewed as an element of GLr+1,soitactsonVΛ.

Proposition 8.8. The action of w0 on VΛ = Vλ is equal to ε(Λ)ηλ,whereε(Λ) = ±1. Therefore, w0(b∂(π))=ε(Λ)b∂(η(π)) for all π ∈ GTΛ (see (8.8)).

Proof. Obviously, the action of w0 on VΛ is compatible with the slr+1-actioninthe following sense: −1 w0(fx)=(w0fw0 )(w0(x)) (8.31) −1 for all f ∈ slr+1,x ∈ VΛ. In view of (7.3), the automorphism f → w0fw0 of slr+1 coincides with the one obtained from the automorphism η of Ur by the specialization q = 1. Comparing (8.31) and (7.5), we see that w0 = ε(Λ)ηλ,whereε(Λ) is the coefficient low of proportionality between w0(1) and bλ . Since both w0 and ηλ are involutions, it follows that ε(Λ) = ±1, and we are done.

28 Corollary 8.9. The number of Ar-tableaux of shape Λ invariant under the Sch¨utzenberger involution η, is equal to |tr (w0,VΛ)|.

The results in this section have interesting combinatorial consequences explored in [11].

References

1. K.Baclawski, A new rule for computing Clebsch-Gordan series, Adv. in Appl. Math., 5 (1984), 416-432. 2. A. Berenstein, A.N.Kirillov, Groups generated by involutions, Gelfand-Tsetlin pat- terns and combinatorics of Young tableaux, RIMS preprint, 1992. 3. A. Berenstein, A. Zelevinsky, Tensor product multiplicities and convex polytopes in partition space, J. Geom.Phys., 5 (1988), 453-472. 4. A. Berenstein, A. Zelevinsky, String bases for quantum groups of type Ar, Advances in Soviet Math., 16, Part 1 (1993), 51-89. 5. E.R.Gansner, On the equality of two plane partition correspondences, Discrete Math., 30 (1980), 121-132. 6. I.M.Gelfand, A.V.Zelevinsky, Polytopes in the pattern space and canonical bases for irreducible representations of gl3, Funct. Anal. Appl., 19, No.2 (1985), 72-75. 7. H.Knight, A.Zelevinsky, Representations of quivers of type A and the multisegment duality, to appear in Advances in Math.. 8. G.Lusztig, Introduction to quantum groups, Birkh¨auser, Boston 1993. 9. M.-P.Sch¨utzenberger, Promotion des morphismes d’ensembles ordonn´es, Discrete Math., 2 (1972), 73-94. 10. J.R.Stembridge, On minuscule representations, plane partitions and involutions in complex Lie groups, Duke Math. J., 73 (1994), 469-490. 11. J.R.Stembridge, Canonical bases and self-evacuating tableaux, preprint, 1994.

29