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

Kazhdan-Lusztig polynomials and subexpressions

Nicolas Libedinsky  and  Geordie Williamson
(Date: July 28, 2025. Preliminary version: NOT FOR DISTRIBUTION; Date: July 28, 2025)
Abstract.

We refine an idea of Deodhar, whose goal is a counting formula for Kazhdan-Lusztig polynomials. This is a consequence of a simple observation that one can use the solution of Soergel’s conjecture to make ambiguities involved in defining certain morphisms between Soergel bimodules in characteristic zero (double leaves) disappear.

1. Introduction

Let (W,S)(W,S) be a Coxeter system. To any pair of elements (x,y)(x,y) of WW, Kazhdan and Lusztig [KL79] associated a polynomial

hx,y[v].h_{x,y}\in{\mathbb{Z}}[v].

These polynomials are ubiquitous in representation theory; they appear in character formulas for simple representations of complex semi-simple Lie algebras, real Lie groups, quantum groups, finite reductive groups …On the other hand, they are still far from being well understood. For example, in several applications the coefficient of vv (the so-called μ\mu-coefficient) plays a crucial role, however even describing when it is non-zero appears extremely subtle.

In their original paper Kazhdan and Lusztig conjectured that the polynomials hx,yh_{x,y} have non-negative coefficients. This conjecture was proved in [KL80] if the underlying Coxeter group is a Weyl or affine Weyl group. The proof proceeds by interpreting hx,yh_{x,y} as the Poincaré polynomial of the local intersection cohomology of a Schubert variety.

Kazhdan and Lusztig’s positivity conjecture was proved in general in [EW14]. The proof is via a study of Soergel bimodules associated to the underlying Coxeter system. Using Soergel bimodules one can produce a space Dx,yD_{x,y} which behaves as though it were the local intersection cohomology of a Schubert variety. The Kazhdan-Lusztig polynomial hx,yh_{x,y} gives the graded dimension of Dx,yD_{x,y}. This implies immediately that hx,yh_{x,y} has non-negative coefficients. The theory also goes quite some way towards explaining what Kazhdan-Lusztig polynomials “are” for arbitrary Coxeter groups.

The aim of this paper is to explain a strategy to use Soergel bimodues to further our combinatorial understanding of Kazhdan-Lusztig polynomials. Our goal (not achieved in this paper) is a “counting formula” for Kazhdan-Lusztig polynomials. Ideally we would like to produce a canonical basis for the space Dx,yD_{x,y}. That is, we would like to find a set Xx,yX_{x,y} and a degree statistic d:Xx,y0d:X_{x,y}\rightarrow{\mathbb{Z}}_{\geq 0} such that if we use Xx,yX_{x,y} and dd to build a positively graded vector space, we have a canonical isomorphism:

eXx,yeDx,y.\bigoplus_{e\in X_{x,y}}{\mathbb{R}}e\stackrel{{\scriptstyle\sim}}{{\rightarrow}}D_{x,y}.

Taking graded dimensions we would deduce a counting formula:

hx,y=eXx,yvd(e).h_{x,y}=\sum_{e\in X_{x,y}}v^{d(e)}.

We expect the sets Xx,yX_{x,y} to reflect in a subtle way the combinatorics of Kazhdan-Lusztig polynomials. If shown to exist, they would open the door to a deeper combinatorial study of Kazhan-Lusztig polynomials.

A proposal for such a counting formula was made by Deodhar in [Deo90]. He considers the set X~x,y¯\widetilde{X}_{x,\underline{y}} of all subexpressions for xx of a fixed reduced expression y¯\underline{y} of yy (see Section 2.1 for more details on our notation). On this set he defines a statistic (“Deodhar’s defect”)

df:X~x,y¯.\operatorname{df}:\widetilde{X}_{x,\underline{y}}\rightarrow{\mathbb{Z}}.

Assuming that Kazhdan-Lusztig polynomials have non-negative coefficients (now known unconditionally), Deodhar proves the existence of a subset Xx,y¯DX~x,y¯X_{x,\underline{y}}^{D}\subset\widetilde{X}_{x,\underline{y}} such that

(1.1) hx,y=eXx,y¯Dvd(e).h_{x,y}=\sum_{e\in X^{D}_{x,\underline{y}}}v^{d(e)}.

Although initially appealing, Deodhar’s proposal suffers from serious drawbacks. The principal one being that the set Xx,y¯DX_{x,\underline{y}}^{D} is not canonical.

There are two sources of non-canonicity. The first is that X~x,y¯\widetilde{X}_{x,\underline{y}} depends on a reduced expression of y¯\underline{y}. We do not regard this dependence as particularly worrisome. Indeed, there are many objects in Lie theory which depend on a choice of reduced expression, and (if canonical up to this point) relating them for different reduced expressions is potentially a fascinating question. The second source of non-canonicity is more concerning: Even for a fixed reduced expression y¯\underline{y} there are in general many possible choices of subsets Xx,y¯DX~x,y¯X_{x,\underline{y}}^{D}\subset\widetilde{X}_{x,\underline{y}} satisfying (1.1). In Deodhar’s framework there is no way to make a distinguished choice. This is as a serious obstacle.

Let x,y¯x,\underline{y} be as above. Using Soergel bimodules one can produce a space Dx,y¯D_{x,\underline{y}} containing Dx,yD_{x,{y}} as a canonical direct summand. In other words, we have a canonical map π:Dx,y¯Dx,y.\pi:D_{x,\underline{y}}\twoheadrightarrow D_{x,{y}}. The following is the main result of this paper.

Theorem 1.1.

There is a canonical isomorphism of graded vector spaces

CLL:eX~x,y¯eDx,y¯.\operatorname{CLL}:\bigoplus_{e\in\widetilde{X}_{x,\underline{y}}}{\mathbb{R}}e\xrightarrow{\sim}D_{x,\underline{y}}.

where the left hand side is graded by Deodhar’s defect, i.e. the generator eX~x,y¯e\in\widetilde{X}_{x,\underline{y}} has degree df(e)\operatorname{df}(e). (CLL stands for “Canonical light leaves”.)

This theorem leads to a natural refinement of Deodhar’s proposal:

Problem 1.2.

Find a subset Xx,y¯LX~x,y¯X_{x,\underline{y}}^{L}\subset\widetilde{X}_{x,\underline{y}} such that the composition of the inclusion, canonical light leaves and the canonical surjection

eXx,y¯LeeX~x,y¯eCLLDx,y¯Dx,y\bigoplus_{e\in X_{x,\underline{y}}^{L}}{\mathbb{R}}e\hookrightarrow\bigoplus_{e\in\widetilde{X}_{x,\underline{y}}}{\mathbb{R}}e\stackrel{{\scriptstyle\operatorname{CLL}}}{{\rightarrow}}D_{x,\underline{y}}\twoheadrightarrow D_{x,y}

is an isomorphism of graded vector spaces.

If the choice of the subset Xx,y¯LX_{x,\underline{y}}^{L} could be made canonically we would regard it as a solution to the counting problem above. Moreover, the map CLL\operatorname{CLL} has the potential to explain why a canonical choice is difficult in general, by recasting the problem as one of linear algebra.

The easiest situation is when the subset of non-zero elements in

{πCLL(e)|eX~x,y¯},\{\pi\circ\operatorname{CLL}(e)\,|\,e\in\widetilde{X}_{x,\underline{y}}\},

already constitutes a basis of Dx,yD_{x,y}. Here we have no choice: we must define Xx,y¯LX_{x,\underline{y}}^{L} to be those ee in X~x,y¯\widetilde{X}_{x,\underline{y}} whose image is non-zero under πCLL\pi\circ\operatorname{CLL}. This situation does occur “in nature”. Namely it is the case for dihedral groups, Universal Coxeter groups, and whenever hx,y=v(y)(x)h_{x,y}=v^{\ell(y)-\ell(x)} (“rationally smooth case”). It is interesting to note that in these cases there already exist closed and combinatorial formulas for Kazhdan-Lusztig polynomials. We feel our result gives a satisfying explanation as to “why” there exist relatively straightforward formulas in these cases.

Remark 1.3.

The basic observation in this paper is that certain morphisms (“light leaves”) may be made canonical in the presence of Soergel’s conjecture. This observation was made during a visit of GW to NL at the Universidad de Chile in 2015, and has been shared with the community since. Subsequently, this idea has been pushed much further: In [Pat20] Patimo studies the case of Grassmannians in detail; and in [LP20] the first author and Patimo study the case of affine type A2A_{2}. In both settings the authors find that the “canonical light leaves”111In the setting of the Grassmannian considered in [Pat20] these are singular variants (in the sense of singular Soergel bimodules) of the maps considered in the present work. associated to different reduced expressions yield many different bases for intersection cohomology, and the question of relating them in interesting ways remains open. In particular, the easy case considered in the previous paragraph is certainly not indicative of the general setting, and the “potentially fascinating question” raised a few paragraphs ago is very much alive. We wrote this paper in order to record the basic observation in the hope that we and others may take it up in the future.

Acknowledgements. The first author was supported by Fondecyt No 1160152.

2. Background

In the following, we recall some standard background in Kazhdan-Lusztig theory and Soergel bimodules. References include [KL79, Soe97, Soe92, Soe07, EW16, Lib08b]. There is also a book [EMTW20] on the way.

2.1. Coxeter group combinatorics

Let (W,S)(W,S) be a Coxeter group with length function \ell and Bruhat order \leq. An expression x¯=(s1,s2,,sm)\underline{x}=(s_{1},s_{2},\dots,s_{m}) is a word in the alphabet SS (i.e. siSs_{i}\in S for all ii). Its length is (x¯)=m\ell(\underline{x})=m.

If x¯=(s1,s2,,sm)\underline{x}=(s_{1},s_{2},\dots,s_{m}) is an expression, we let x:=s1s2smx:=s_{1}s_{2}\dots s_{m} denote the product in WW. Given an expression x¯=(s1,s2,,sm)\underline{x}=(s_{1},s_{2},\dots,s_{m}), a subexpression of x¯\underline{x} is a word e¯=e1e2em\underline{e}=e_{1}e_{2}\dots e_{m} of length mm in the alphabet {0,1}\{0,1\}. We will write e¯x¯\underline{e}\subset\underline{x} to indicate that e¯\underline{e} is a subexpression of x¯\underline{x}. We set

x¯e¯:=s1e1s2e2smemW\underline{x}^{\underline{e}}:=s_{1}^{e_{1}}s_{2}^{e_{2}}\dots s_{m}^{e_{m}}\in W

and say that e¯x¯\underline{e}\subset\underline{x} expresses x¯e¯\underline{x}^{\underline{e}}.

For 1im1\leq i\leq m, we define wi:=s1e1s2e2sieiw_{i}:=s_{1}^{e_{1}}s_{2}^{e_{2}}\dots s_{i}^{e_{i}}. We also define di{U,D}d_{i}\in\{U,D\} (where UU stands for Up and DD for Down) in the following way:

di:={Uif wi1si>wi1,Dif wi1si<wi1.d_{i}:=\begin{cases}U&\text{if }w_{i-1}s_{i}>w_{i-1},\\ D&\text{if }w_{i-1}s_{i}<w_{i-1}.\end{cases}

We write the decorated sequence (d1e1,,dmem)(d_{1}e_{1},\ldots,d_{m}e_{m}). Deodhar’s defect df\operatorname{df} is defined by

df(e):=|{i|diei=U0}||{i|diei=D0}|\operatorname{df}(e):=|\{\,i\,|\,d_{i}e_{i}=U0\}|-|\{\,i\,|\,d_{i}e_{i}=D0\}|

2.2. Hecke algebras

For the basic definitions of Hecke algebras and Kazhdan-Lusztig polynomials we follow [Soe97]. Let (W,S)(W,S) be a Coxeter system. Recall that the Hecke algebra \mathcal{H} of (W,S)(W,S) is the algebra with free [v,v1]\mathbb{Z}[v,v^{-1}]-basis given by symbols {hx}xW\{h_{x}\}_{x\in W} and multiplication given by

hxhs:={hxsif xs>x,(v1v)hx+hxsif xs<x.h_{x}h_{s}:=\begin{cases}h_{xs}&\text{if }xs>x,\\ (v^{-1}-v)h_{x}+h_{xs}&\text{if }xs<x.\end{cases}

We can define a \mathbb{Z}-module morphism ()¯:\overline{(-)}:\mathcal{H}\rightarrow\mathcal{H} by the formula v¯=v1\overline{v}=v^{-1} and hx¯=(hx1)1.\overline{h_{x}}=(h_{x^{-1}})^{-1}. It is a ring morphism, and we call it the duality in the Hecke algebra. The Kazhdan-Lusztig basis of \mathcal{H} is denoted by {bx}xW\{b_{x}\}_{x\in W}. It is a [v,v1]\mathbb{Z}[v,v^{-1}]-basis of \mathcal{H} and it is characterised by the two conditions

bx¯=bxandbxhx+yWv[v]hy\overline{b_{x}}=b_{x}\qquad\text{and}\qquad b_{x}\in h_{x}+\sum_{y\in W}v\mathbb{Z}[v]h_{y}

for all xWx\in W. If we write bx=hx+yWhy,xhyb_{x}=h_{x}+\sum_{y\in W}h_{y,x}h_{y} then the Kazhdan-Lusztig polynomials (as defined in [KL79]) py,xp_{y,x} are defined by the formula py,x=vl(x)l(y)hy,x,p_{y,x}=v^{l(x)-l(y)}h_{y,x}, and Cx=bxC^{\prime}_{x}=b_{x} (their q1/2q^{-1/2} is our vv).

Let us define the [v,v1]\mathbb{Z}[v,v^{-1}]-bilinear form

(,):×[v,v1],(-,-):\mathcal{H}\times\mathcal{H}\rightarrow\mathbb{Z}[v,v^{-1}],

given by (hx,hy):=δx,y(h_{x},h_{y}):=\delta_{x,y}. A useful property of this pairing is that (bx,by)v[v](b_{x},b_{y})\in v\mathbb{Z}[v] if xyx\neq y and (bx,bx)1+v[v](b_{x},b_{x})\in 1+v\mathbb{Z}[v].

2.3. Soergel bimodules

We fix a realisation 𝔥{\mathfrak{h}} of our Coxeter system (W,S)(W,S) over the real numbers {\mathbb{R}}. That is, 𝔥{\mathfrak{h}} is a real vector space and we have fixed roots {αs}sS𝔥\{\alpha_{s}\}_{s\in S}\subset{\mathfrak{h}}^{*} and coroots {αs}sS𝔥\{\alpha_{s}^{\vee}\}_{s\in S}\subset{\mathfrak{h}} such that the familiar formulas from Lie theory define a representation of WW of 𝔥{\mathfrak{h}} and 𝔥{\mathfrak{h}}^{*}.

Throughout, we assume that this is a realisation for which Soergel’s conjecture holds. For example we could take 𝔥{\mathfrak{h}} to be the realisation from [Soe07, EW14]. We could also take 𝔥{\mathfrak{h}} to be the geometric representation [Lib08a] so that 𝔥=αs{\mathfrak{h}}=\bigoplus{\mathbb{R}}\alpha_{s}^{\vee} and for tS,t\in S, the element αt𝔥\alpha_{t}\in{\mathfrak{h}}^{*} is defined by αt,αs=cos(π/mst)\langle\alpha_{t},\alpha_{s}^{\vee}\rangle=-\cos(\pi/m_{st}), where mstm_{st} denotes the order (possibly \infty) of stWst\in W.

Having fixed 𝔥{\mathfrak{h}} we define R=S(𝔥)=𝒪(𝔥)R=S({\mathfrak{h}}^{*})=\mathcal{O}({\mathfrak{h}}) to be the symmetric algebra on 𝔥{\mathfrak{h}}^{*} (alias the polynomials functions on 𝔥{\mathfrak{h}}), graded so that 𝔥{\mathfrak{h}}^{*} has degree 2. We denote by BimR\textrm{Bim}_{R} the category of \mathbb{Z}-graded RR-bimodules which are finitely generated both as left and right RR-modules. Given an object M=MiBimRM=\bigoplus M^{i}\in\textrm{Bim}_{R} we denote by M(k)M(k) the shifted bimodule, with M(k)i:=Mk+iM(k)^{i}:=M^{k+i}. Given objects M,NBimRM,N\in\textrm{Bim}_{R} we denote their tensor product by juxtaposition: MN:=MRNMN:=M\otimes_{R}N. This operation makes BimR\textrm{Bim}_{R} into a monoidal category. The Krull-Schmidt theorem holds in BimR\textrm{Bim}_{R}.

For any sSs\in S we denote by RsRR^{s}\subset R the ss-invariants in RR. We consider the bimodule

Bs:=RRsR(1).B_{s}:=R\otimes_{R^{s}}R(1).

Given an expression w¯=(s1,,sm)\underline{w}=(s_{1},\dots,s_{m}) we consider the Bott-Samelson bimodule

Bw¯:=Bs1Bs2Bsm.B_{\underline{w}}:=B_{s_{1}}B_{s_{2}}\dots B_{s_{m}}.

The category {\mathcal{B}} of Soergel bimodules is defined to be the full, strict (i.e. closed under isomorphism), additive (i.e. M,NMNM,N\in{\mathcal{B}}\Rightarrow M\oplus N\in{\mathcal{B}}), monoidal (i.e. M,NMNM,N\in{\mathcal{B}}\Rightarrow MN\in{\mathcal{B}}) category of BimR\textrm{Bim}_{R} which contains BsB_{s} for all sSs\in S and is closed under shifts (m)(m) and direct summands.

Notation 2.1.

For Soergel bimodules MM and NN, we denote by Homi(M,N)\mathrm{Hom}^{i}(M,N) the degree ii morphisms in Hom(M,N),\mathrm{Hom}(M,N), where the latter is the set of all RR-bimodule morphisms.

2.4. Soergel’s theorems and Soergel’s conjecture

Soergel proved the following facts (usually known as Soergel’s categorification theorem). For all wWw\in W there exists a unique (up to isomorphism) bimodule BwB_{w} which occurs as a direct summand of Bw¯B_{\underline{w}} for any reduced expression w¯\underline{w} of ww, and is not a summand of (some shift of) By¯B_{\underline{y}} for any shorter sequence y¯\underline{y}. The set {Bw|wW}\{B_{w}\,|\,w\in W\} constitutes a complete set of non-isomorphic indecomposable Soergel bimodules, up to isomorphism and grading shift. There is a unique isomorphism of [v,v1]\mathbb{Z}[v,v^{-1}]-algebras between the split Grothendieck group of {\mathcal{B}} and the Hecke algebra

ch:[],\mathrm{ch}:[{\mathcal{B}}]\rightarrow\mathcal{H},

satisfying ch([Bs])=bs\mathrm{ch}([B_{s}])=b_{s} and ch([R(1)])=v\mathrm{ch}([R(1)])=v.

Soergel gave a formula to calculate the graded dimensions of the Hom spaces in {\mathcal{B}} in the Hecke algebra. We need some notation to explain it. Given a finite dimensional graded \mathbb{R}-vector space V=ViV=\oplus V^{i}, we define

gdim(V)=dim(Vi)vi0[v,v1].{\mathrm{gdim}}(V)=\sum\mathrm{dim}(V^{i})v^{i}\in{\mathbb{Z}}_{\geq 0}[v,v^{-1}].

Given a finitely-generated and free graded right RR-module MM, we define

grk(M):=gdim(MR).{\mathrm{grk}}(M):={\mathrm{gdim}}(M\otimes_{R}\mathbb{R}).

The following is Soergel’s hom formula. Let M,NM,N\in{\mathcal{B}}, then Hom(M,N)\operatorname{Hom}(M,N) is finitely-generated and free as a right RR-module, and

grkHom(M,N)=(ch(M),ch(N)¯).{\mathrm{grk}}\mathrm{Hom}(M,N)={(\mathrm{ch}(M),\overline{\mathrm{ch}(N)})}.

Soergel’s conjecture (now a theorem by Elias and the second author [EW14]) is the following statement:

ch([Bx])=bxfor all xW.\mathrm{ch}([B_{x}])=b_{x}\quad\text{for all $x\in W.$}

We remark that when Soergel’s conjecture is satisfied (the case considered in this paper), by Soergel’s hom formula and by the useful property at the end of Section 2.2, we obtain a complete description of the degree zero morphisms between indecomposable objects:

(2.1) Hom0(Bx,By)δx,y.\mathrm{Hom}_{0}(B_{x},B_{y})\cong\delta_{x,y}\mathbb{R}.

2.5. Double leaves

An important result in the theory of Soergel bimodules is a theorem of the first author giving a “double leaves” basis of morphisms between Soergel bimodules. Let w¯=(s1,,sm)\underline{w}=(s_{1},\dots,s_{m}) denote an expression. For any subexpression e¯\underline{e} of w¯\underline{w} the first author associates a morphism

LLw¯,e¯:Bw¯Bx¯(df(e¯)).\operatorname{LL}_{\underline{w},\underline{e}}:B_{\underline{w}}\rightarrow B_{\underline{x}}(\operatorname{df}(\underline{e})).

Here x¯\underline{x} is a fixed but arbitrary reduced expression of x=w¯e¯x=\underline{w}^{\underline{e}}. The definition of LLw¯,e¯\operatorname{LL}_{\underline{w},\underline{e}} is inductive, and will not be given here, as we will not need it. However it is important to note that the definition of LLw¯,e¯\operatorname{LL}_{\underline{w},\underline{e}} depends on choices (fixed reduced expressions for elements and fixed sequences of braid relations between reduced expressions) which seem difficult to make canonical.

However, once one has fixed such choices one can produce a basis of homomorphisms between any two Bott-Samelson bimodules. Indeed, a theorem of the first author [Lib15, Thm. 3.2] (see also [EW16, Thm 6.11]) asserts that the set

xW{LLw¯,e¯LLz¯,f¯|e¯w¯,f¯z¯ such that w¯e¯=z¯f¯=x}\bigsqcup_{x\in W}\{\operatorname{LL}_{\underline{w},\underline{e}}^{*}\circ\operatorname{LL}_{\underline{z},\underline{f}}\;|\;\underline{e}\subset\underline{w},\underline{f}\subset\underline{z}\text{ such that }\underline{w}^{\underline{e}}=\underline{z}^{\underline{f}}=x\}

is a free RR-basis for Hom(Bz¯,Bw¯)\mathrm{Hom}(B_{\underline{z}},B_{\underline{w}}).

2.6. The sets Dx,y¯D_{x,\underline{y}} and Dx,yD_{x,y}

Let M,NM,N\in{\mathcal{B}}. For xWx\in W we denote by

Hom<x(M,N)Hom(M,N)\operatorname{\mathrm{Hom}}_{<x}(M,N)\subset\operatorname{\mathrm{Hom}}(M,N)

the vector space generated by all morphisms f:MNf:M\rightarrow N that factor through By(n)B_{y}(n) for some y<xy<x and nn\in\mathbb{Z}. Let

Homx(M,N):=Hom(M,N)/Hom<x(M,N).\operatorname{\mathrm{Hom}}_{\not<x}(M,N):=\operatorname{\mathrm{Hom}}(M,N)/\operatorname{\mathrm{Hom}}_{<x}(M,N).

We denote by x{\mathcal{B}}_{\not<x} the category whose objects coincide with those of {\mathcal{B}} and for any M,NxM,N\in{\mathcal{B}}_{\not<x} we have Homx(M,N):=Homx(M,N)\operatorname{\mathrm{Hom}}_{{\mathcal{B}}_{\not<x}}(M,N):=\operatorname{\mathrm{Hom}}_{\not<x}(M,N).

Consider the sets

D^x,y¯:=Homx(By¯,Bx),\widehat{D}_{x,\underline{y}}:=\operatorname{\mathrm{Hom}}_{\not<x}(B_{\underline{y}},B_{x}),
Dx,y¯:=Homx(By¯,Bx)RandD_{x,\underline{y}}:=\operatorname{\mathrm{Hom}}_{\not<x}(B_{\underline{y}},B_{x})\otimes_{R}\mathbb{R}\ \mathrm{and}
Dx,y:=Homx(By,Bx)R.D_{x,y}:=\operatorname{\mathrm{Hom}}_{\not<x}(B_{{y}},B_{x})\otimes_{R}\mathbb{R}.

The set Dx,yD_{x,y} is a canonical direct summand of Dx,y¯D_{x,\underline{y}}. This is because, when Soergel’s conjecture is satisfied, there is one element in End(By¯)\mathrm{End}(B_{\underline{y}}) projecting to ByB_{y} called the favorite projector (see [Lib15, §4.1]). Let us give the construction of this projector. Let us assume (by induction) that projection and inclusion maps have been constructed

By¯py¯Byiy¯By¯B_{\underline{y}}\stackrel{{\scriptstyle p_{\underline{y}}}}{{\twoheadrightarrow}}B_{y}\stackrel{{\scriptstyle i_{\underline{y}}}}{{\hookrightarrow}}B_{\underline{y}}

for some reduced expression y¯\underline{y} of yy. Suppose y<ys,y<ys, then

bybs=bys+x<ysmxbx,withmx0.b_{y}b_{s}=b_{ys}+\sum_{x<ys}m_{x}b_{x},\ \ \mathrm{with}\ m_{x}\in\mathbb{Z}_{\geq 0}.

By Soergel’s conjecture this implies

ByBs=Bysx<ysBxmx.B_{y}B_{s}=B_{ys}\oplus\bigoplus_{x<ys}B_{x}^{\oplus m_{x}}.

By (2.1), there is only one projector in this space projecting to BysB_{ys}, which we write as

ByBspy,sBysiy,sByBs.B_{y}B_{s}\stackrel{{\scriptstyle p_{y,s}}}{{\twoheadrightarrow}}B_{ys}\stackrel{{\scriptstyle i_{y,s}}}{{\hookrightarrow}}B_{y}B_{s}.

We now define the inclusion and projection maps of our favourite projector to be the compositions

By¯Bspy¯idBsByBspy,sBysiy,sByBsiy¯idBsBy¯Bs.B_{\underline{y}}B_{s}\stackrel{{\scriptstyle p_{\underline{y}}\operatorname{id}_{B_{s}}}}{{\twoheadrightarrow}}B_{y}B_{s}\stackrel{{\scriptstyle p_{y,s}}}{{\twoheadrightarrow}}B_{ys}\stackrel{{\scriptstyle i_{y,s}}}{{\hookrightarrow}}B_{y}B_{s}\stackrel{{\scriptstyle i_{\underline{y}}\operatorname{id}_{B_{s}}}}{{\hookrightarrow}}B_{\underline{y}}B_{s}.

3. Canonical light leaves

This section contains the new observations of this paper. We explain that certain canonical elements and maps allow one to define canonical light leaves, from which our main theorem (Theorem 1.1) follows easily.

Remark 3.1.

In this paper we use “canonical” to mean “not depending on any choices”. We do not use it in the stronger sense that is typical in Lie theory (i.e. to refer to the Kazhdan-Lusztig basis of the Hecke algebra, or the canonical basis of quantum groups).

3.1. Some canonical elements

What do we really mean when we write BxB_{x}? In the general setting of Soergel bimodules, we mean a representative of an equivalence class of isomorphic bimodules, where each isomorphism is not canonical. In our setting (where Soergel’s conjecture is available), we mean a representative of an equivalence class of isomorphic bimodules, where each isomorphism is canonical up to an invertible scalar (in our case {\mathbb{R}}^{*}). We now explain a somewhat adhoc way to fix this scalar, so that BxB_{x} is defined up to unique isomorphism.

Consider an expression x¯\underline{x}, and the corresponding Bott-Samelson bimodule Bx¯B_{\underline{x}}. It contains a canonical element

cbotx¯:=111Bx¯.c_{\textrm{bot}}^{\underline{x}}:=1\otimes 1\otimes\dots\otimes 1\in B_{\underline{x}}.

(Note that Bx¯B_{\underline{x}} is zero below degree (x¯)-\ell(\underline{x}) and is spanned by cbotc_{\textrm{bot}} in degree (x¯)-\ell(\underline{x}); bot stands for “bottom”.) We denote by cbotxBxc_{\textrm{bot}}^{x}\in B_{x} the image of cbotx¯c_{\textrm{bot}}^{\underline{x}} under the favourite projector, where x¯\underline{x} is a reduced expression for xx.

From now on we will always understand BxB_{x} to mean BxB_{x} together with the element cbotBxc_{\textrm{bot}}\in B_{x}. Given two representatives (Bx,cbotx)(B_{x},c_{\textrm{bot}}^{x}) and (B~x,c~botx)(\tilde{B}_{x},\tilde{c}_{\textrm{bot}}^{x}), there is a unique isomorphism BxB~xB_{x}\rightarrow\tilde{B}_{x} which sends cbotxc_{\textrm{bot}}^{x} to c~botx\tilde{c}_{\textrm{bot}}^{x}.

Remark 3.2.

Consider the following commutative diagram

Bx¯B_{\underline{x}}Bx¯B_{\underline{x}^{\prime}}BxB_{x}BxB_{x}φ\varphipx¯p_{\underline{x}}px¯p_{\underline{x}^{\prime}}\simζ\zeta

where: φ\varphi is a braid move (see [EW16, §4.2], where they are called rex moves); px¯p_{\underline{x}} (resp. px¯p_{\underline{x}^{\prime}}) are the projections in the favourite projector associated to x¯\underline{x} and x¯\underline{x}^{\prime}; and ζ\zeta is the induced isomorphism. One may check that ζ(cbotx)=cbotx\zeta(c_{\textrm{bot}}^{x})=c_{\textrm{bot}}^{x}. (We will not need this fact below.) This gives another sense to which cbotc_{\textrm{bot}} is canonical.

3.2. Some canonical maps

In this section we introduce the canonical maps which will be our building blocks for the definition of canonical light leaves, in the next section.

Lemma 3.3.

Let xWx\in W and sSs\in S and suppose that x<xsx<xs. The spaces

Hom0(BxBs,Bxs),Hom1(BxsBs,Bxs)andHom1(Bxs,Bx)\mathrm{Hom}^{0}(B_{x}B_{s},B_{xs}),\quad\mathrm{Hom}^{-1}(B_{xs}B_{s},B_{xs})\quad\text{and}\quad\mathrm{Hom}^{1}(B_{xs},B_{x})

are all one-dimensional.

Proof.

We consider the spaces one at a time. As in last section, we have

BxBs=Bxsy<xsBymyB_{x}B_{s}=B_{xs}\oplus\bigoplus_{y<xs}B_{y}^{\oplus m_{y}}

and (2.1) allows us to conclude that Hom0(BxBs,Bxs)\mathrm{Hom}^{0}(B_{x}B_{s},B_{xs}) is one dimensional.

We now consider the second space. By Soergel’s hom formula and Soergel’s conjecture, the dimension of

Hom1(BxsBs,Bxs)\mathrm{Hom}^{-1}(B_{xs}B_{s},B_{xs})

is the coefficient of v1v^{-1} in the Laurent polynomial (bxsbs,bxs)(b_{xs}b_{s},b_{xs}). But

bxsbs=(v+v1)bxs.b_{xs}b_{s}=(v+v^{-1})b_{xs}.

As (bxs,bxs)1+v[v](b_{xs},b_{xs})\in 1+v\mathbb{Z}[v], we conclude that Hom1(BxsBs,Bxs).\mathrm{Hom}^{-1}(B_{xs}B_{s},B_{xs})\cong\mathbb{R}.

For the last case, we need to calculate the coefficient of vv in (bxs,bx)(b_{xs},b_{x}), i.e. in

(hxs+vhx+y<xsyxPyhy,hx+z<xQzhz)(h_{xs}+vh_{x}+\sum_{\begin{subarray}{c}y<xs\\ y\neq x\end{subarray}}P_{y}h_{y}\ ,\ h_{x}+\sum_{z<x}Q_{z}h_{z})

where Py,Qzv[v]P_{y},Q_{z}\in v\mathbb{Z}[v]. By definition of the pairing, it is clear that the coefficient of vv is 11. ∎

Let xWx\in W and sSs\in S be as in the lemma above (i.e. x<xsx<xs). Both BxBsB_{x}B_{s} and BxsB_{xs} are one-dimensional in degree (x)1-\ell(x)-1, where they are spanned by cbotxcbotsc_{\textrm{bot}}^{x}c_{\textrm{bot}}^{s} and cbotxsc_{\textrm{bot}}^{xs} respectively. (We write cbotxcbotsc_{\textrm{bot}}^{x}c_{\textrm{bot}}^{s} instead of cbotxcbotsc_{\textrm{bot}}^{x}\otimes c_{\textrm{bot}}^{s}.) Hence there exists a unique map

(3.1) αx,s:BxBsBxs\alpha_{x,s}:B_{x}B_{s}\rightarrow B_{xs}

which maps cbotxcbotsc_{\textrm{bot}}^{x}c_{\textrm{bot}}^{s} to cbotxsc_{\textrm{bot}}^{xs}. Similar considerations show that there exists a unique map

(3.2) βx,s:BxsBsBxs(1)\beta_{x,s}:B_{xs}B_{s}\rightarrow B_{xs}(1)

resp.

(3.3) γx,s:BxsBx(1)\gamma_{x,s}:B_{xs}\rightarrow B_{x}(1)

mapping cbotxscbotsc_{\textrm{bot}}^{xs}c_{\textrm{bot}}^{s} to cbotxsc_{\textrm{bot}}^{xs} (resp. cbotxsc_{\textrm{bot}}^{xs} to cbotxc_{\textrm{bot}}^{x}).

3.3. The construction

We will use the maps αx,s,βx,s\alpha_{x,s},\beta_{x,s} and γx,s\gamma_{x,s} constructed above. We will also use the multiplication map

ms:BsR(1):fgfg.m_{s}:B_{s}\rightarrow R(1):f\otimes g\mapsto fg.
Remark 3.4.

The reader may easily check that in fact ms=γid,sm_{s}=\gamma_{\operatorname{id},s}.

Consider the following data:

  1. (1)

    an expression (not necessarily reduced) y¯=(s1,,sn)\underline{y}=(s_{1},\ldots,s_{n});

  2. (2)

    elements xWx\in W, sSs\in S; and

  3. (3)

    f:By¯Bxf:B_{\underline{y}}\rightarrow B_{x}.

To this data, we will associate two new maps:

f0:By¯BsBxandf1:By¯BsBxs.f0:B_{\underline{y}}B_{s}\rightarrow B_{x}\quad\text{and}\quad f1:B_{\underline{y}}B_{s}\rightarrow B_{xs}.

These maps are constructed as follows: If x<xsx<xs, define

f0:=fmsandf1:=αx,s(fid).f0:=f\otimes m_{s}\quad\text{and}\quad f1:=\alpha_{x,s}\circ(f\otimes\mathrm{id}).

If xs<xxs<x, define

f0:=βxs,s(fid)andf1:=γxs,sβxs,s(fid).f0:=\beta_{xs,s}\circ(f\otimes\mathrm{id})\quad\text{and}\quad f1:=\gamma_{xs,s}\circ\beta_{xs,s}\circ(f\otimes\mathrm{id}).

Given an expression w¯\underline{w} and a subexpression e¯\underline{e} define the canonical light leaf

CLLw¯,e¯:=ide¯,\operatorname{CLL}_{\underline{w},\underline{e}}:=\mathrm{id}\underline{e},

where id\mathrm{id} means idEnd(R)\mathrm{id}\in\mathrm{End}(R) and for example id(0,1,0)\mathrm{id}(0,1,0) means (((id0)1)0)(((\mathrm{id}0)1)0).

Example 3.5.

If x¯=(s1,,sm)\underline{x}=(s_{1},\dots,s_{m}) is reduced, and e¯=(1,1,,1)\underline{e}=(1,1,\dots,1) then CLLw¯,e¯\operatorname{CLL}_{\underline{w},\underline{e}} agrees with the projection in the favourite projector. If e¯=(0,0,,0)\underline{e}=(0,0,\dots,0) then CLLw¯,e¯=ms1msm\operatorname{CLL}_{\underline{w},\underline{e}}=m_{s_{1}}\otimes\dots\otimes m_{s_{m}}.

The proof of the following theorem is essentially the same as in [Lib15, Thm. 3.2] and [EW16, Thm 6.11].

Theorem 3.6.

The set

xW{CLLw¯,e¯CLLz¯,f¯|e¯w¯,f¯z¯ such that w¯e¯=z¯f¯=x}\bigsqcup_{x\in W}\{\operatorname{CLL}_{\underline{w},\underline{e}}^{*}\circ\operatorname{CLL}_{\underline{z},\underline{f}}\;|\;\underline{e}\subset\underline{w},\underline{f}\subset\underline{z}\text{ such that }\underline{w}^{\underline{e}}=\underline{z}^{\underline{f}}=x\}

is a free RR-basis for Hom(Bz¯,Bw¯)\mathrm{Hom}(B_{\underline{z}},B_{\underline{w}}).

Now we can explain why this theorem proves Theorem 1.1. By Theorem 3.6, the graded set {CLLy¯,e¯|withe¯expressingx}\{\operatorname{CLL}_{\underline{y},\underline{e}}\,|\,\mathrm{with}\,\underline{e}\,\mathrm{expressing}\,x\} is naturally an RR-basis of D^x,y¯\widehat{D}_{x,\underline{y}}, thus it gives an \mathbb{R}-basis of Dx,y¯.D_{x,\underline{y}}. So, in summary, the canonical map CLL in Theorem 1.1 is the \mathbb{R}-linear map defined on the generators e¯X~x,y¯\underline{e}\in\widetilde{X}_{x,\underline{y}} by e¯CLLy¯,e¯.\underline{e}\mapsto\mathrm{CLL}_{\underline{y},\underline{e}}.

References

  • [Deo90] V. V. Deodhar. A combinatorial setting for questions in Kazhdan-Lusztig theory. Geom. Dedicata, 36(1):95–119, 1990.
  • [EMTW20] B. Elias, S. Makisumi, U. Thiel, and G. Williamson. Introduction to Soergel bimodules. RSME series. Springer, Berlin, 2020. to appear.
  • [EW14] B. Elias and G. Williamson. The Hodge theory of Soergel bimodules. Ann. of Math. (2), 180(3):1089–1136, 2014.
  • [EW16] B. Elias and G. Williamson. Soergel calculus. Represent. Theory, 20:295–374, 2016.
  • [KL79] D. Kazhdan and G. Lusztig. Representations of Coxeter groups and Hecke algebras. Invent. Math., 53(2):165–184, 1979.
  • [KL80] D. Kazhdan and G. Lusztig. Schubert varieties and Poincaré duality. In Geometry of the Laplace operator (Proc. Sympos. Pure Math., Univ. Hawaii, Honolulu, Hawaii, 1979), Proc. Sympos. Pure Math., XXXVI, pages 185–203. Amer. Math. Soc., Providence, R.I., 1980.
  • [Lib08a] N. Libedinsky. Équivalences entre conjectures de Soergel. J. Algebra, 320(7):2695–2705, 2008.
  • [Lib08b] N. Libedinsky. Sur la catégorie des bimodules de Soergel. J. Algebra, 320(7):2675–2694, 2008.
  • [Lib15] N. Libedinsky. Light leaves and Lusztig’s conjecture. Adv. Math., 280:772–807, 2015.
  • [LP20] N. Libedinsky and L. Patimo. On the affine hecke category. Preprint, 2020.
  • [Pat20] L. Patimo. Bases of the intersection cohomology of grassmannian schubert varieties. Preprint, 2020.
  • [Soe92] W. Soergel. The combinatorics of Harish-Chandra bimodules. J. Reine Angew. Math., 429:49–74, 1992.
  • [Soe97] W. Soergel. Kazhdan-Lusztig polynomials and a combinatoric[s] for tilting modules. Represent. Theory, 1:83–114 (electronic), 1997.
  • [Soe07] W. Soergel. Kazhdan-Lusztig-Polynome und unzerlegbare Bimoduln über Polynomringen. J. Inst. Math. Jussieu, 6(3):501–525, 2007.