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

Equivariant K-theory and Tangent spaces to Schubert varieties

William Graham Department of Mathematics, University of Georgia, Boyd Graduate Studies Research Center, Athens, GA 30602 [email protected]  and  Victor Kreiman Department of Mathematics, University of Wisconsin - Parkside, Kenosha, WI 53141 [email protected]
Abstract.

Tangent spaces to Schubert varieties of type AA were characterized by Lakshmibai and Seshadri [LS84]. This result was extended to the other classical types by Lakshmibai [Lak95], [Lak00b], and [Lak00a]. We give a uniform characterization of tangent spaces to Schubert varieties in cominuscule G/PG/P. Our results extend beyond cominuscule G/PG/P; they describe the tangent space to any Schubert variety in G/BG/B at a point xBxB, where xx is a cominuscule Weyl group element in the sense of Peterson. Our results also give partial information about the tangent space to any Schubert variety at any point. Our method is to describe the tangent spaces of Kazhdan-Lusztig varieties, and then recover results for Schubert varieties. Our proof uses a relationship between weights of the tangent space of a variety with torus action, and factors of the class of the variety in torus equivariant KK-theory. The proof relies on a formula for Schubert classes in equivariant KK-theory due to Graham [Gra02] and Willems [Wil06], as well as a theorem on subword complexes due to Knutson and Miller [KM04], [KM05].

1991 Mathematics Subject Classification:
Primary 14M15; Secondary 05E14. Keywords: K-theory, Schubert variety, flag variety, Grassmannian, cominuscule

1. Introduction

One goal in the study of Schubert varieties is to understand their singularities. A related goal is to understand their Zariski tangent spaces, or equivalently, the weights of their Zariski tangent spaces at fixed points of the action of a maximal torus. A description of these tangent spaces in type A was given by Lakshmibai and Seshadri [LS84]. In [Lak95], [Lak00b], and [Lak00a], Lakshmibai extended this result to all classical types (see also [BL00, Chapter 5]). We give a different description of tangent spaces to Schubert varieties, which is uniform across all types. Our description, however, recovers only part of the tangent space, except in certain cases, such as Schubert varieties in cominuscule G/PG/P, in which it recovers the entire tangent space. Our results hold for an algebraically closed ground field kk of characteristic 0. The results of [LS84], [Lak95], [Lak00b], and [Lak00a] hold in arbitrary characteristic.

Rather than studying Schubert varieties directly, we focus on the smaller Kazhdan-Lusztig varieties, which differ locally only by a well-prescribed affine space. We study general Kazhdan-Lusztig varieties, but we do not attempt to recover all weights of the tangent space. Rather, we restrict our attention to those weights of the tangent space which are integrally indecomposable in an ambient space VV, which is to say that they cannot be written as the sum of other weights of VV. We characterize such weights. When all weights of VV are integrally indecomposable in VV, our characterization captures all weights of the tangent space. This occurs, for example, for Kazhdan-Lusztig varieties in cominuscule G/PG/P, or more generally, for any Kazhdan-Lusztig variety at a TT-fixed point (i.e., point of tangency) which is a cominuscule Weyl group element.

1.1. Statement of results

Let GG be a semisimple algebraic group defined over an algebraically closed field kk of characteristic 0. Let PBTP\supseteq B\supseteq T be a parabolic subgroup, Borel subgroup, and maximal torus of GG respectively. We denote the set of weights of a representation EE of TT by Φ(E)\Phi(E). Let WW be the Weyl group of (G,T)(G,T), and SS the set of simple reflections in WW relative to BB.

Fix wxWw\leq x\in W. Let XwX^{w} be the Schubert variety BwB¯\overline{B^{-}wB}, and YxwY_{x}^{w} the Kazhdan-Lusztig variety BxBBwB¯BxB\cap\overline{B^{-}wB}, in G/BG/B. The Kazhdan-Lusztig variety YxwY_{x}^{w} (and thus its tangent space at xx, TxYxwT_{x}Y_{x}^{w}) is an affine subvariety of an ambient space VV in G/BG/B with weights Φ(V)=I(x1)\Phi(V)=I(x^{-1}), the inversion set of x1x^{-1}. If 𝐬=(s1,,sl)\mathbf{s}=(s_{1},\ldots,s_{l}), siSs_{i}\in S, is a reduced expression for xx, then the elements of I(x1)I(x^{-1}) are given explicitly by the formula γi=s1si1(αi)\gamma_{i}=s_{1}\cdots s_{i-1}(\alpha_{i}), i=1,,li=1,\ldots,l, where αi\alpha_{i} is the simple root corresponding to sis_{i}.

Our main result is the following theorem (see Theorem 5.8):

Theorem A. Suppose γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}). Then the following are equivalent:

  1. (i)

    γjΦ(TxYxw)\gamma_{j}\in\Phi(T_{x}Y_{x}^{w}).

  2. (ii)

    There exists a reduced subexpression of (s1,,s^j,,sl)(s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}) for ww.

  3. (iii)

    The Demazure product of (s1,,s^j,,sl)(s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}) is greater than or equal to ww.

This theorem, which applies to Kazhdan-Lusztig varieties in G/BG/B, extends to Schubert varieties and to G/PG/P. Moreover, when xx is a cominuscule Weyl group element of WW, all γj\gamma_{j} are integrally indecomposable in I(x1)I(x^{-1}), so Theorem A recovers all weights of the tangent space.

Remark 1.1.

If γj\gamma_{j} is not integrally indecomposable in I(x1)I(x^{-1}), then (ii) and (iii) of Theorem A are still equivalent, but (i) is no longer equivalent to (ii) and (iii) in general.

Remark 1.2.

Let us denote by TExYxwTE_{x}Y_{x}^{w} the span of the tangent lines to TT-invariant curves through xx in YxwY_{x}^{w}; then TExYxwTxYxwTE_{x}Y_{x}^{w}\subseteq T_{x}Y_{x}^{w}. It is known that condition (iii) of Theorem A, with the Demazure product of (s1,,s^j,,sl)(s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}) replaced by the ordinary product s1s^jsls_{1}\cdots\widehat{s}_{j}\cdots s_{l}, gives a characterization of all weights of TExYxwTE_{x}Y_{x}^{w} (and not just the integrally indecomposable weights) [Car95] [CK03]. Thus, Theorem A can be viewed as a characterization of the integrally indecomposable weights of TxYxwT_{x}Y_{x}^{w} which is similar to this known characterization of all weights of the smaller space TExYxwTE_{x}Y_{x}^{w}.

Remark 1.3.

The paper [GK21] proves that in simply-laced types, the Demazure product of (s1,,s^j,,sl)(s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}) of Theorem A (iii) is equal to the ordinary product s1s^jsls_{1}\cdots\widehat{s}_{j}\cdots s_{l}, provided that γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}). As a corollary, it is proved that in simply-laced types, when xx is a cominuscule Weyl group element, Φ(TxXw)=Φ(TExXw)\Phi(T_{x}X^{w})=\Phi(TE_{x}X^{w}).

1.2. Outline of proof

Our proof of Theorem A uses equivariant KK-theory. Let us fix notation and give some basic definitions and properties. If TT acts on a smooth scheme MM, the Grothendieck group of TT-equivariant coherent sheaves (or vector bundles) on MM is denoted by KT(M)K_{T}(M). If NN is a TT-stable subscheme of MM, then the class in KT(M)K_{T}(M) of the pushfoward of the structure sheaf 𝒪N\mathcal{O}_{N} of NN is denoted by [𝒪N]M[\mathcal{O}_{N}]_{M}, or sometimes just [𝒪N][\mathcal{O}_{N}]. A TT-equivariant vector bundle on a point is a representation of TT, so KT({point})K_{T}(\{\text{point}\}) can be identified with R(T)R(T), the representation ring of TT. The inclusion im:{m}Mi_{m}:\{m\}\to M of a TT-fixed point induces a pullback im:KT(M)KT({m})=R(T)i_{m}^{*}:K_{T}(M)\to K_{T}(\{m\})=R(T).

Consider for the moment a more general situation than that of the previous subsection: VV a representation of TT such that all weights of VV lie in an open half-space and have multiplicity one, YVY\subseteq V a TT-stable subscheme, and xYx\in Y a TT-fixed point. The structure sheaf 𝒪Y\mathcal{O}_{Y} defines a class [𝒪Y]KT(V)[\mathcal{O}_{Y}]\in K_{T}(V). We show that the factors of ix[𝒪Y]R(T)i_{x}^{*}[\mathcal{O}_{Y}]\in R(T) contain information about the tangent space TxYT_{x}Y. Let us say that 1eθ1-e^{-\theta} is a simple factor of ix[𝒪Y]i_{x}^{*}[\mathcal{O}_{Y}] if ix[𝒪Y]=(1eθ)Qi_{x}^{*}[\mathcal{O}_{Y}]=(1-e^{-\theta})Q for some QR(T)Q\in R(T) which is a polynomial in eλe^{-\lambda}, λΦ(V){θ}\lambda\in\Phi(V)\setminus\{\theta\}. We prove (see Proposition 3.5)

Proposition B. Suppose θ\theta is integrally indecomposable in Φ(V)\Phi(V). Then θΦ(TxY)\theta\in\Phi(T_{x}Y) if and only if 1eθ1-e^{-\theta} is not a simple factor of ix[𝒪Y]i_{x}^{*}[\mathcal{O}_{Y}].

Now set VV and xx as in Subsection 1.1 and set YY to be the Kazhdan-Lusztig variety YxwY_{x}^{w}. For θΦ(V)=I(x1)\theta\in\Phi(V)=I(x^{-1}), we have θ=γj\theta=\gamma_{j} for some jj. Proposition B then becomes

Proposition C. Suppose γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}). Then γjΦ(TxYxw)\gamma_{j}\in\Phi(T_{x}Y_{x}^{w}) if and only if 1eγj1-e^{-\gamma_{j}} is not a simple factor of ix[𝒪Yxw]i_{x}^{*}[\mathcal{O}_{Y_{x}^{w}}].

This characterization of Φ(TxYxw)\Phi(T_{x}Y_{x}^{w}) would appear to suffer from a computational difficulty: determining whether 1eγj1-e^{-\gamma_{j}} is a factor of ix[𝒪Yxw]i_{x}^{*}[\mathcal{O}_{Y_{x}^{w}}], let alone whether it is a simple factor, seems to be a nontrivial problem. It requires some sort of division algorithm in R(T)R(T). One approach would be to search for an expression for ix[𝒪Yxw]i_{x}^{*}[\mathcal{O}_{Y_{x}^{w}}] as a sum of terms in which 1eγj1-e^{-\gamma_{j}} appears explicitly as a factor of each summand. To rule out the possibility that 1eγj1-e^{-\gamma_{j}} is a factor of ix[𝒪Yxw]i_{x}^{*}[\mathcal{O}_{Y_{x}^{w}}], then, one would need to show that no such expression exists. This would presumably require knowledge of all possible expressions for ix[𝒪Yxw]i_{x}^{*}[\mathcal{O}_{Y_{x}^{w}}].

We show a way around this apparent computational difficulty. When γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}), the question of whether or not 1eγj1-e^{-\gamma_{j}} is a simple factor of ix[𝒪Yxw]i_{x}^{*}[\mathcal{O}_{Y_{x}^{w}}] can be answered by using a single expression for ix[𝒪Yxw]i_{x}^{*}[\mathcal{O}_{Y_{x}^{w}}] due to Graham [Gra02] and Willems [Wil06]:

ix[𝒪Yxw]=𝐭𝒯w,𝐬(1)e(𝐭)i𝐭(1eγi)R(T),i_{x}^{*}[\mathcal{O}_{Y_{x}^{w}}]=\sum_{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}}(-1)^{e(\mathbf{t})}\prod_{i\in\mathbf{t}}(1-e^{-\gamma_{i}})\in R(T), (1.1)

where 𝒯w,𝐬\mathcal{T}_{w,\mathbf{s}} is the set of sequences 𝐭=(i1,,im)\mathbf{t}=(i_{1},\ldots,i_{m}), 1i1<<iml1\leq i_{1}<\cdots<i_{m}\leq l, such that Hsi1Hsim=HwH_{s_{i_{1}}}\cdots H_{s_{i_{m}}}=H_{w} in the 0-Hecke algebra, and e(𝐭)=m(w)e(\mathbf{t})=m-\ell(w). More precisely, we prove (see Theorem 5.6):

Theorem D. Suppose γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}). Then 1eγj1-e^{-\gamma_{j}} is a simple factor of ix[𝒪Yxw]i_{x}^{*}[\mathcal{O}_{Y_{x}^{w}}] if and only if 1eγj1-e^{-\gamma_{j}} occurs explicitly as a factor of every summand of 𝐭𝒯w,𝐬(1)e(𝐭)i𝐭(1eγi)\sum_{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}}(-1)^{e(\mathbf{t})}\prod_{i\in\mathbf{t}}(1-e^{-\gamma_{i}}), i.e., if and only if jj belongs to every 𝐭𝒯w,𝐬\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}.

We note that one direction of this theorem follows immediately from (1.1). Combining Proposition C and Theorem D yields

Theorem E. Suppose that γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}). Then γjΦ(TxYxw)\gamma_{j}\in\Phi(T_{x}Y_{x}^{w}) if and only if jj does not belong to every 𝐭𝒯w,𝐬\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}.

Now the equivalence of (i) and (ii) of Theorem A is essentially a reformulation of Theorem E, using some properties of 0-Hecke algebras. The equivalence of (ii) and (iii) is due to Knutson-Miller [KM04, Lemma 3.4 (1)].

The paper is organized as follows. In Section 2, we recall definitions and properties of equivariant KK-theory and weights of tangent spaces to schemes with TT-actions. In Section 3 we prove Proposition B. In Section 4, we give a corollary to a theorem by Knutson-Miller on subword complexes [KM04], [KM05]. Our proof of Theorem D relies on this corollary. In Section 5, we apply the material of the previous sections to Kazhdan-Lusztig varieties in order to prove Proposition C and Theorems D and E. In Section 6, we show how to extend these results to G/PG/P and discuss the case of cominuscule Weyl group elements and cominuscule G/PG/P.

The related paper [GK21] examines rationally indecomposable weights of the ambient space VV. Rational indecomposability is a stricter condition than integral indecomposability, so the set of rationally indecomposable weights is contained in the set of integrally indecomposable weights. For this smaller set of weights, [GK21] obtains stronger results. For example, it is shown that the elements of Φ(TxYxw)\Phi(T_{x}Y_{x}^{w}) which are rationally indecomposable in I(x1)I(x^{-1}) lie in Φ(TExYxw)\Phi(TE_{x}Y_{x}^{w}). Several results of [GK21] rely on those of this paper.

2. Preliminaries

Let kk be an algebraically closed field of characteristic 0. We work in the category of schemes over kk. A point on a scheme will always refer to a closed point. If WW is a finite dimensional vector space over kk, then one can give a bijection between the vectors of WW and the (closed) points of the affine scheme Spec(Sym(W))\operatorname{Spec}(\operatorname{Sym}(W^{*})) [GW10, Corollary 1.11]. As is customary, we will often identify WW with this scheme, and in this context refer to Sym(W)\operatorname{Sym}(W^{*}) as the coordinate ring of WW.

In this section, we collect information concerning equivariant KK-theory, tangent spaces, and tangent cones. We include proofs for the convenience of the reader.

2.1. The pullback to a fixed point in TT-equivariant KK-theory

Let T=(k)nT=(k^{*})^{n} be a torus, and let T^=Hom(T,k)\widehat{T}=\operatorname{Hom}(T,k^{*}) be the character group of TT. The mapping λdλ\lambda\mapsto d\lambda from a character to its differential at 1T1\in T embeds T^\widehat{T} in the dual 𝔱{\mathfrak{t}}^{*} of the Lie algebra of TT. We will usually view T^\widehat{T} as a subset of 𝔱{\mathfrak{t}}^{*} under this embedding and express the group operation additively. If λ\lambda denotes an element of T^\widehat{T} viewed as an element of 𝔱{\mathfrak{t}}^{*}, then the corresponding homomorphism TkT\to k^{*} is written as eλe^{\lambda}. The representation ring R(T)R(T) is defined to be the free \mathbb{Z}-module with basis eλe^{\lambda}, λT^\lambda\in\widehat{T}, with multiplication given by eλeμ=eλ+μe^{\lambda}e^{\mu}=e^{\lambda+\mu}.

Let VV be a finite dimensional representation of TT such that all weights of VV have multiplicity one and lie in an open half-space in the real span of the characters of TT. Denote the set of weights of TT on VV by Φ(V)\Phi(V) and the set of nonnegative integer linear combinations of elements of Φ(V)\Phi(V) in 𝔱{\mathfrak{t}}^{*} by ConeΦ(V)\operatorname{Cone}_{\mathbb{Z}}\Phi(V). The dual representation VV^{*} has weights Φ(V)-\Phi(V), and the coordinate ring k[V]=Sym(V)k[V]=\operatorname{Sym}(V^{*}) of VV has weights ConeΦ(V)-\operatorname{Cone}_{\mathbb{Z}}\Phi(V). Denote λΦ(V)(1eλ)\prod_{\lambda\in\Phi(V)}(1-e^{-\lambda}) by λ1(V)\lambda_{-1}(V^{*}). For ΦAΦ(V)\Phi_{A}\subseteq\Phi(V), let [eλ,λΦA]\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi_{A}}] denote the subring of R(T)R(T) generated over \mathbb{Z} by eλe^{-\lambda}, λΦA\lambda\in\Phi_{A}.

Lemma 2.1.

If νConeΦ(V)\nu\in\operatorname{Cone}_{\mathbb{Z}}\Phi(V), then eνe^{-\nu} can be expressed as a monomial in eλe^{-\lambda}, λΦ(V)\lambda\in\Phi(V).

Proof.

Write ν=c1λ1++ctλt\nu=c_{1}\lambda_{1}+\cdots+c_{t}\lambda_{t}, where λiΦ(V)\lambda_{i}\in\Phi(V) and cic_{i} are nonnegative integers. Then eν=eciλi=(eλi)cie^{-\nu}=e^{\sum-c_{i}\lambda_{i}}=\prod(e^{-\lambda_{i}})^{c_{i}}. ∎

The map i0:KT(V)R(T)i_{0}^{*}:K_{T}(V)\to R(T) is an isomorphism, which we denote by iVi_{V}^{*}.

Lemma 2.2.

Let YY be a TT-stable closed subscheme of VV. Then iV[𝒪Y]V[eλ,λΦ(V)]i_{V}^{*}[\mathcal{O}_{Y}]_{V}\in\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi(V)}].

Proof.

We adapt a method appearing in [Ros89]. Denote k[V]k[V] by RR. All modules in this proof will be TT-stable RR-modules, and all maps TT-equivariant RR-homomorphisms.

Consider the projection

Rk[Y]0R\to k[Y]\to 0 (2.1)

The kernel is a TT-stable ideal II of RR which is generated by a finite number of weight vectors r1,1,,r1,n1r_{1,1},\ldots,r_{1,n_{1}}. Let ν1,j\nu_{1,j} be the weight of r1,jr_{1,j}; kν1,jk_{\nu_{1,j}} the TT-representation of weight ν1,j\nu_{1,j}; and F1=j=0n1Rkν1,jF_{1}=\oplus_{j=0}^{n_{1}}R\otimes k_{\nu_{1,j}}. Note that RR acts on the first factor of F1F_{1} and TT on both, and that Φ(F1)Φ(R)\Phi(F_{1})\subseteq\Phi(R). There exists a map f1:F1Rf_{1}:F_{1}\to R such that

F1Rk[Y]0F_{1}\to R\to k[Y]\to 0 (2.2)

is exact (f1f_{1} maps 111\otimes 1 from the jj-th summand of F1F_{1} to r1,jr_{1,j}).

The kernel of f1f_{1} is finitely generated over RR (since F1F_{1} is finitely generated and RR is Noetherian), and thus is generated by a finite number of weight vectors. Thus the above procedure can be repeated to produce a module F2F_{2} and map F2F1F_{2}\to F_{1}, which, when appended to (2.2), yields an exact sequence. Moreover, Φ(F2)Φ(F1)Φ(R)\Phi(F_{2})\subseteq\Phi(F_{1})\subseteq\Phi(R). When iterated, this procedure must terminate, by the Hilbert Syzygy Theorem. The resulting complex is a resolution of k[Y]k[Y]:

0FdF1Rk[Y]00\to F_{d}\to\cdots\to F_{1}\to R\to k[Y]\to 0 (2.3)

where Fi=jRkνi,jF_{i}=\oplus_{j}R\otimes k_{\nu_{i,j}} and Φ(Fi)Φ(R)\Phi(F_{i})\subseteq\Phi(R). Thus νi,jΦ(Fi)Φ(R)=ConeΦ(V)\nu_{i,j}\in\Phi(F_{i})\subseteq\Phi(R)=-\operatorname{Cone}_{\mathbb{Z}}\Phi(V).

The resolution (2.3) corresponds to a resolution of 𝒪Y\mathcal{O}_{Y} over 𝒪V\mathcal{O}_{V}:

0d1𝒪V𝒪Y00\to\mathcal{F}_{d}\to\cdots\to\mathcal{F}_{1}\to\mathcal{O}_{V}\to\mathcal{O}_{Y}\to 0

where i=j𝒪Vkνi,j\mathcal{F}_{i}=\oplus_{j}\mathcal{O}_{V}\otimes k_{\nu_{i,j}}. Since [𝒪V]V=1[\mathcal{O}_{V}]_{V}=1, we have

iV[𝒪Y]V=1+i,j(1)ieνi,jiV[𝒪V]V=1+i,j(1)ieνi,ji_{V}^{*}[\mathcal{O}_{Y}]_{V}=1+\sum_{i,j}(-1)^{i}e^{\nu_{i,j}}i_{V}^{*}[\mathcal{O}_{V}]_{V}=1+\sum_{i,j}(-1)^{i}e^{\nu_{i,j}}

By Lemma 2.1, this lies in [eλ,λΦ(V)]\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi(V)}]. ∎

2.2. Weights of tangent and normal spaces

The coordinate ring of VV is Sym(V)=k[xλ,λΦ(V)]\operatorname{Sym}(V^{*})=k[x_{\lambda},\lambda\in\Phi(V)], a polynomial ring, where xλx_{\lambda} denotes a vector of VV^{*} of weight λ-\lambda. Observe that this polynomial ring is graded by the coordinates xλx_{\lambda} and also by the weights of the TT action. The weights of any TT-stable subspace of VV form a subset of Φ(V)\Phi(V), whose corresponding weight vectors span the subspace. Thus there is a bijection between the TT-stable subspaces of VV and the subsets of Φ(V)\Phi(V). The coordinate ring of a TT-stable subspace ZZ is Sym(Z)=k[xλ,λΦ(Z)]\operatorname{Sym}(Z^{*})=k[x_{\lambda},\lambda\in\Phi(Z)].

Let YVY\to V be a TT-equivariant closed immersion, with TT-fixed point xYx\in Y mapping to 0V0\in V. Let 𝔪{\mathfrak{m}} be the maximal ideal of the local ring of YY at xx. The tangent space to YY at xx, denoted TxY{T_{x}Y}, is defined to be (𝔪/𝔪2)({\mathfrak{m}}/{\mathfrak{m}}^{2})^{*}, a vector space over kk. It embeds naturally in the tangent space to VV at 0, which is isomorphic to VV [GW10, (6.2), (6.3)].

Let

B=Sym(𝔪/𝔪2),C=i0𝔪i/𝔪i+1.B=\operatorname{Sym}({\mathfrak{m}}/{\mathfrak{m}}^{2}),\qquad C=\oplus_{i\geq 0}{\mathfrak{m}}^{i}/{\mathfrak{m}}^{i+1}.

Note that the degree one components of BB and CC, which are denoted by B1B_{1} and C1C_{1} respectively, are both equal to 𝔪/𝔪2{\mathfrak{m}}/{\mathfrak{m}}^{2}. We will often identify TxY{T_{x}Y} with the affine scheme Spec(B)\operatorname{Spec}(B). The tangent cone to YY at xx, denoted TCxYTC_{x}Y, is defined to be Spec(C)\operatorname{Spec}(C). The projection BCB\twoheadrightarrow C induces an inclusion TCxYTxY{TC_{x}Y}\hookrightarrow{T_{x}Y}.

Both TxY{T_{x}Y} and TCxY{TC_{x}Y} are TT-stable, and TCxYTxY{TC_{x}Y}\hookrightarrow{T_{x}Y} is TT-equivariant. The coordinate ring BB of TxY{T_{x}Y} is equal to k[xλ,λΦ(TxY)]k[x_{\lambda},\lambda\in\Phi({T_{x}Y})], with character

CharB=1λΦ(TxY)(1eλ)=1λ1((TxY)).\operatorname{Char}B=\frac{1}{\prod_{\lambda\in\Phi({T_{x}Y})}(1-e^{-\lambda})}=\frac{1}{\lambda_{-1}(({T_{x}Y})^{*})}.

This lives in R^(T)\widehat{R}(T), the set of expressions of the form λT^cλeλ\sum_{\lambda\in\widehat{T}}c_{\lambda}e^{\lambda}. Similarly, we have a formula for the character of CC.

Proposition 2.3.

CharC=iV[𝒪TCxY]Vλ1(V)=iTxY[𝒪TCxY]TxYλ1((TxY))\operatorname{Char}C=\dfrac{i_{V}^{*}[\mathcal{O}_{{TC_{x}Y}}]_{V}}{\lambda_{-1}(V^{*})}=\dfrac{i_{T_{x}Y}^{*}[\mathcal{O}_{{TC_{x}Y}}]_{T_{x}Y}}{\lambda_{-1}(({T_{x}Y})^{*})}.

Proof.

The first equality is proved in [GK15, Proposition 2.1] and the second in [GK17, (3.10)]. ∎

Proposition 2.4.

[𝒪TCxY]V=[𝒪Y]V[\mathcal{O}_{{TC_{x}Y}}]_{V}=[\mathcal{O}_{Y}]_{V} in KT(V)K_{T}(V).

Proof.

See [GK17, Proposition 3.1(2)]. ∎

3. Factors in equivariant KK-theory

We keep the notations, definitions, and assumptions of the previous section. Denote Φ(V),Φ(TxY)\Phi(V),\Phi({T_{x}Y}), and Φ(V/TxY)=Φ(V)Φ(TxY)\Phi(V/{T_{x}Y})=\Phi(V)\setminus\Phi({T_{x}Y}) by Φamb,Φtan\Phi_{\operatorname{amb}},\Phi_{\tan}, and Φnor\Phi_{\operatorname{nor}} respectively. Then Φamb=ΦtanΦnor\Phi_{\operatorname{amb}}=\Phi_{\tan}\sqcup\Phi_{\operatorname{nor}}. If P[eλ,λΦamb]P\in\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi_{\operatorname{amb}}}], then we will say that 1eθ1-e^{-\theta} is a simple factor of PP if P=(1eθ)QP=(1-e^{-\theta})Q, for Q[eλ,λΦamb{θ}]Q\in\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi_{\operatorname{amb}}\setminus\{\theta\}}].

Example 3.1.

Suppose λ1,λ2,λ3,λ4Φamb\lambda_{1},\lambda_{2},\lambda_{3},\lambda_{4}\in\Phi_{\operatorname{amb}} are distinct, and λ3=λ1+λ2\lambda_{3}=\lambda_{1}+\lambda_{2}. Consider

P=(1eλ1)(1eλ4)+(1eλ2)(1eλ4)(1eλ1)(1eλ2)(1eλ4),P=(1-e^{-\lambda_{1}})(1-e^{-\lambda_{4}})+(1-e^{-\lambda_{2}})(1-e^{-\lambda_{4}})-(1-e^{-\lambda_{1}})(1-e^{-\lambda_{2}})(1-e^{-\lambda_{4}}),

an element of [eλ,λΦamb]\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi_{\operatorname{amb}}}]. Then PP can be expressed as (1e(λ1+λ2))(1eλ4)(1-e^{-(\lambda_{1}+\lambda_{2})})(1-e^{-\lambda_{4}}). Thus both 1e(λ1+λ2)1-e^{-(\lambda_{1}+\lambda_{2})} and 1eλ41-e^{-\lambda_{4}} are simple factors of PP.

Remark 3.2.

In Section 5.3, we will need to distinguish between the nature of the two factors 1e(λ1+λ2)1-e^{-(\lambda_{1}+\lambda_{2})} and 1eλ41-e^{-\lambda_{4}} of PP in Example 3.1. While the latter factor appears explicitly as a factor of each summand, and thus is easily identifiable as a factor of PP, the former does not. We will refer to 1eλ41-e^{-\lambda_{4}} as an explicit factor of the expression PP.

There are usually many ways to express an element P[eλ,λΦamb]P\in\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi_{\operatorname{amb}}}]. Explicit factors depend on the particular expression of PP, while (non-explicit) factors do not.

We wish to study whether it is possible to determine whether a weight θ\theta lies in Φnor\Phi_{\operatorname{nor}} or Φtan\Phi_{\tan} based on whether or not 1eθ1-e^{-\theta} is a simple factor of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V}. We begin with the following observation:

Proposition 3.3.

If θΦnor\theta\in\Phi_{\operatorname{nor}}, then 1eθ1-e^{-\theta} is a simple factor of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V}.

Proof.

By Propositions 2.4 and 2.3,

iV[𝒪Y]V=iV[𝒪TCxY]V=λ1(V)λ1((TxY))iTxY[𝒪TCxY]TxY.i_{V}^{*}[\mathcal{O}_{Y}]_{V}=i_{V}^{*}[\mathcal{O}_{{TC_{x}Y}}]_{V}=\frac{\lambda_{-1}(V^{*})}{\lambda_{-1}(({T_{x}Y})^{*})}i_{T_{x}Y}^{*}[\mathcal{O}_{{TC_{x}Y}}]_{T_{x}Y}.

Now, λ1(V)λ1((TxY))=λ1((V/TxY))=λΦnor(1eλ)\frac{\lambda_{-1}(V^{*})}{\lambda_{-1}(({T_{x}Y})^{*})}=\lambda_{-1}((V/{T_{x}Y})^{*})=\prod_{\lambda\in\Phi_{\operatorname{nor}}}(1-e^{-\lambda}), and 1eθ1-e^{-\theta} occurs among the terms of this product exactly once. Moreover, since TCxY{TC_{x}Y} is a closed algebraic subscheme of TxY{T_{x}Y}, Lemma 2.2 implies iTxY[𝒪TCxY]TxY[eλ,λΦ(TxY)][eλ,λΦamb{θ}]i_{T_{x}Y}^{*}[\mathcal{O}_{{TC_{x}Y}}]_{T_{x}Y}\in\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi({T_{x}Y})}]\subseteq\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi_{\operatorname{amb}}\setminus\{\theta\}}]. ∎

The converse of this proposition is false, as illustrated by the following example.

Example 3.4.

Suppose that TT acts on V=k3V=k^{3}, and that the standard basis vectors 𝐞1,𝐞2,𝐞3\mathbf{e}_{1},\mathbf{e}_{2},\mathbf{e}_{3} are weight vectors with corresponding weights λ1,λ2,λ3=λ1+λ2\lambda_{1},\lambda_{2},\lambda_{3}=\lambda_{1}+\lambda_{2}. Letting x1,x2,x3Vx_{1},x_{2},x_{3}\in V^{*} denote the dual of the standard basis, we have, for tTt\in T, txi=eλi(t)xitx_{i}=e^{-\lambda_{i}}(t)x_{i}, i=1,2,3i=1,2,3. Let YY be the affine subscheme of VV defined by the ideal I=(x1x2)I=(x_{1}x_{2}), and let xx be the origin. Then the ideal of the tangent space and tangent cone of YY at xx are {0}\{0\} and (x1x2)(x_{1}x_{2}) respectively. The tangent space of YY at xx is all of VV, so Φtan=Φamb\Phi_{\tan}=\Phi_{\operatorname{amb}} and Φnor\Phi_{\operatorname{nor}} is empty. The tangent cone of YY at xx is the union of the x2x3x_{2}x_{3}-plane and the x1x3x_{1}x_{3} plane, so its coordinate ring CC has character

CharC=(11eλ1+11eλ21)11e(λ1+λ2)\operatorname{Char}C=\left(\frac{1}{1-e^{-\lambda_{1}}}+\frac{1}{1-e^{-\lambda_{2}}}-1\right)\frac{1}{1-e^{-(\lambda_{1}+\lambda_{2})}}

Additionally, λ1(V)=(1eλ1)(1eλ2)(1e(λ1+λ2))\lambda_{-1}(V^{*})=(1-e^{-\lambda_{1}})(1-e^{-\lambda_{2}})(1-e^{-(\lambda_{1}+\lambda_{2})}), and thus by Proposition 2.3,

iV[𝒪Y]V=iV[𝒪C]V\displaystyle i_{V}^{*}[\mathcal{O}_{Y}]_{V}=i_{V}^{*}[\mathcal{O}_{C}]_{V} =(1eλ1)+(1eλ2)(1eλ1)(1eλ2)\displaystyle=(1-e^{-\lambda_{1}})+(1-e^{-\lambda_{2}})-(1-e^{-\lambda_{1}})(1-e^{-\lambda_{2}})
=1e(λ1+λ2)\displaystyle=1-e^{-(\lambda_{1}+\lambda_{2})}

Hence 1eλ3=1e(λ1+λ2)1-e^{-\lambda_{3}}=1-e^{-(\lambda_{1}+\lambda_{2})} is a simple factor of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V}, but λ3\lambda_{3} lies in Φtan\Phi_{\tan}.

In this example, the fact that there exists λ3\lambda_{3} in Φtan\Phi_{\tan} such that 1eλ31-e^{-\lambda_{3}} is a simple factor of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V}, thus violating the converse of Proposition 3.3, appears to be related to the fact that λ3\lambda_{3} can be expressed as the sum of other weights of Φamb\Phi_{\operatorname{amb}}. This suggests that the converse of Proposition 3.3 may hold if we restrict to weights θ\theta which cannot be expressed as such a sum. This assertion is true, and is proved in the following proposition. Let us say that a weight of Φamb\Phi_{\operatorname{amb}} is integrally decomposable if it can be expressed as a positive integer linear combination of other elements of Φamb\Phi_{\operatorname{amb}}, or integrally indecomposable otherwise.

Proposition 3.5.

Let θ\theta be an integrally indecomposable element of Φamb\Phi_{\operatorname{amb}}. Then the following are equivalent:

  1. (i)

    θΦtan\theta\in\Phi_{\tan}.

  2. (ii)

    xθB1x_{\theta}\in B_{1}.

  3. (iii)

    xθC1x_{\theta}\in C_{1}.

  4. (iv)

    θ-\theta is a weight of CC.

  5. (v)

    1eθ1-e^{-\theta} is not a simple factor of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V}.

Proof.

(v) \Rightarrow (i) by Proposition 3.3; (i) \Rightarrow (ii) since B=k[xλ,λΦtan]B=k[x_{\lambda},\lambda\in\Phi_{\tan}]; (ii) \Rightarrow (iii) since C1=𝔪/𝔪2=B1C_{1}={\mathfrak{m}}/{\mathfrak{m}}^{2}=B_{1}; and (iii) \Rightarrow (iv) because xθx_{\theta} has weight θ-\theta.

It remains to prove (iv) \Rightarrow (v). We prove the contrapositive. Thus assume that 1eθ1-e^{-\theta} is a simple factor of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V}. Then iV[𝒪Y]V=(1eθ)Qi_{V}^{*}[\mathcal{O}_{Y}]_{V}=(1-e^{-\theta})Q, where Q[eλ,λΦamb{θ}]Q\in\mathbb{Z}[e^{-\lambda},\lambda\in{\Phi_{\operatorname{amb}}\setminus\{\theta\}}]. By Propositions 2.3 and 2.4,

CharC=QλΦamb{θ}(1eλ)=QλΦamb{θ}(1+eλ+e2λ+).\operatorname{Char}C=\frac{Q}{\prod_{\lambda\in\Phi_{\operatorname{amb}}\setminus\{\theta\}}(1-e^{-\lambda})}=Q\prod\nolimits_{\lambda\in\Phi_{\operatorname{amb}}\setminus\{\theta\}}(1+e^{-\lambda}+e^{-2\lambda}+\cdots).

Expanding, one obtains an infinite sum of terms eνe^{-\nu}, νCone(Φamb)\nu\in\operatorname{Cone}_{\mathbb{Z}}(\Phi_{\operatorname{amb}}). None of these terms is equal to eθe^{-\theta}. (This is because none of the factors in the above product for CharC\operatorname{Char}C contain a term eθe^{-\theta}; since θ\theta is integrally indecomposable in Φamb\Phi_{\operatorname{amb}}, the term eθe^{-\theta} cannot be obtained by expanding the product.) Thus θ-\theta is not a weight of CC, as required. ∎

4. Euler characteristics of subword complexes

In Section 5, we will apply the results of the previous section to Schubert varieties and Kazhdan-Lusztig varieties. One main tool for this purpose is Corollary 4.6, whose proof relies on a theorem by Knutson-Miller on subword complexes [KM05], [KM04].

4.1. The reduced Euler characteristic

In this subsection we give a brief review of simplicial complexes and their Euler characteristics.

Recall that an (abstract) simplicial complex on a finite set AA is a set Δ\Delta of subsets of AA, called faces, with the property that if FΔF\in\Delta and GFG\subseteq F then GΔG\in\Delta. The dimension of a face FF is #F1\#F-1, and the dimension of Δ\Delta is the maximum dimension of a face. A maximal face of Δ\Delta is called a facet. Note that Δ=\Delta=\emptyset and Δ={}\Delta=\{\emptyset\} are distinct simplicial complexes, called the void complex and irrelevant complex respectively. If Δ\Delta\neq\emptyset, then \emptyset must be a face of Δ\Delta.

The reduced Euler characteristic of Δ\Delta is defined to be χ~(Δ)=FΔ(1)dimF\widetilde{\chi}(\Delta)=\sum_{F\in\Delta}(-1)^{\dim F}. If Δ\Delta\neq\emptyset, so that Δ\emptyset\in\Delta, then \emptyset contributes a summand of 1-1 to χ~(Δ)\widetilde{\chi}(\Delta). From this we see, for example, that χ~({})=1\widetilde{\chi}(\{\emptyset\})=-1, but χ~()=0\widetilde{\chi}(\emptyset)=0.

Suppose that Δ\Delta\neq\emptyset or {}\{\emptyset\}. Denoting the elements of AA by x1,,xmx_{1},\ldots,x_{m}, the set AA can be embedded in m\mathbb{R}^{m} by mapping xix_{i} to the iith standard basis vector of m\mathbb{R}^{m}. For any face FF of Δ\Delta, let |F||F| be the convex hull of its vertices in m\mathbb{R}^{m}. The geometric realization of Δ\Delta is then defined to be |Δ|=FΔ|F||\Delta|=\bigcup_{F\in\Delta}|F|, a topological subspace of m\mathbb{R}^{m}. If |Δ||\Delta| is homeomorphic to a topological space YY, then Δ\Delta is called a triangulation of YY. In this case, the reduced Euler characteristic of Δ\Delta is equal to the topological reduced Euler characteristic of YY. If YY is a manifold with boundary and its boundary Y\partial Y is nonempty, then there exists a subcomplex of Δ\Delta which is a triangulation of Y\partial Y [Mau80, Proposition 5.4.4]. This subcomplex is called the boundary of Δ\Delta and denoted by Δ\partial\Delta.

For m0m\geq 0, let Bm={xm:x1}B^{m}=\{x\in\mathbb{R}^{m}:\|x\|\leq 1\} and Sm={xm+1:x=1}S^{m}=\{x\in\mathbb{R}^{m+1}:\|x\|=1\}, the mm-ball and mm-sphere respectively. Both can be triangulated. In the sequel, when we refer to an mm-ball or mm-sphere or their notations, m0m\geq 0, we will mean a triangulation of the object. When we refer to the sphere S1S^{-1}, we will mean the irrelevant complex {}\{\emptyset\}. With these conventions, for m0m\geq 0, χ~(Bm)=0\widetilde{\chi}(B^{m})=0, χ~(Sm1)=(1)m1\widetilde{\chi}(S^{m-1})=(-1)^{m-1}, Bm=Sm1\partial B^{m}=S^{m-1}, and Sm1=\partial S^{m-1}=\emptyset. (Observe that B0\partial B^{0} is the irrelevant complex, but Sm1\partial S^{m-1} is the void complex.) For Δ=Bm\Delta=B^{m} or SmS^{m}, define χ~(Δ)=χ~(Δ)χ~(Δ)\widetilde{\chi}^{\circ}(\Delta)=\widetilde{\chi}(\Delta)-\widetilde{\chi}(\partial\Delta).

Lemma 4.1.

χ~(Bm)=χ~(Sm)=(1)m\widetilde{\chi}^{\circ}(B^{m})=\widetilde{\chi}^{\circ}(S^{m})=(-1)^{m}, for m0m\geq 0.

Proof.

The proof is a calculation: χ~(Bm)=0(1)m1=(1)m\widetilde{\chi}^{\circ}(B^{m})=0-(-1)^{m-1}=(-1)^{m}, and χ~(Sm)=(1)m0=(1)m\widetilde{\chi}^{\circ}(S^{m})=(-1)^{m}-0=(-1)^{m}. ∎

4.2. 0-Hecke algebras

Let GG be a semisimple algebraic group, BB a Borel subgroup, BB^{-} the opposite Borel subgroup, and T=BBT=B\cap B^{-} a maximal torus. Let W=NG(T)/TW=N_{G}(T)/T, the Weyl group of GG. Let SS be the set of simple reflections of WW relative to BB. The 0-Hecke algebra \mathcal{H} associated to (W,S)(W,S) over a commutative ring RR is the associative RR-algebra generated by HuH_{u}, uWu\in W, and subject to the following relations: H1H_{1} is the identity element, and if uWu\in W and sSs\in S, then HuHs=HusH_{u}H_{s}=H_{us} if (us)>(u)\ell(us)>\ell(u) and HuHs=HuH_{u}H_{s}=H_{u} if (us)<(u)\ell(us)<\ell(u). If 𝐪=(q1,,ql)\mathbf{q}=(q_{1},\ldots,q_{l}) is any sequence of elements of SS, define the Demazure product δ(𝐪)W\delta(\mathbf{q})\in W by the equation Hq1Hql=Hδ(𝐪)H_{q_{1}}\cdots H_{q_{l}}=H_{\delta(\mathbf{q})}. Define (𝐪)=l\ell(\mathbf{q})=l and e(𝐪)=(𝐪)(δ(𝐪))e(\mathbf{q})=\ell(\mathbf{q})-\ell(\delta(\mathbf{q})).

4.3. Subword Complexes

Let 𝐬=(s1,,sl)\mathbf{s}=(s_{1},\ldots,s_{l}) be a sequence of elements of SS and wWw\in W. The subword complex Δ(𝐬,w)\Delta(\mathbf{s},w) is defined to be the set of subsequences 𝐫=(si1,,sit)\mathbf{r}=(s_{i_{1}},\ldots,s_{i_{t}}), 1i1<<itl1\leq i_{1}<\cdots<i_{t}\leq l, whose complementary subsequence 𝐬𝐫\mathbf{s}\setminus\mathbf{r} contains a reduced expression for ww. One checks that Δ(𝐬,w)\Delta(\mathbf{s},w) is a simplicial complex. Subword complexes were introduced in [KM04], [KM05]. We will require that 𝐬\mathbf{s} contains a reduced expression for ww.

Remark 4.2.

The requirement that 𝐬\mathbf{s} contains a reduced expression for ww implies that the empty sequence \emptyset lies in Δ(𝐬,w)\Delta(\mathbf{s},w). In particular, Δ(𝐬,w)\Delta(\mathbf{s},w) is not the void complex. It is possible, however, for Δ(𝐬,w)\Delta(\mathbf{s},w) to be the irrelevant complex. For example, this occurs when 𝐬=(s1)\mathbf{s}=(s_{1}), w=s1w=s_{1}.

The following theorem is [KM04, Theorem 3.7]:

Theorem 4.3.

The subword complex Δ(𝐬,w)\Delta(\mathbf{s},w) is either a ball or sphere. A face 𝐫\mathbf{r} is in the boundary of Δ(𝐬,w)\Delta(\mathbf{s},w) if and only if δ(𝐬𝐫)w\delta(\mathbf{s}\setminus\mathbf{r})\neq w.

Corollary 4.4.

𝐪𝐬δ(𝐪)=w(1)e(𝐪)=1\displaystyle\sum_{\mathbf{q}\subseteq\mathbf{s}\atop\delta(\mathbf{q})=w}(-1)^{e(\mathbf{q})}=1.

Proof.

If 𝐫𝐬\mathbf{r}\subseteq\mathbf{s}, then 𝐫Δ(𝐬,w)\mathbf{r}\in\Delta(\mathbf{s},w) if and only if 𝐬𝐫\mathbf{s}\setminus\mathbf{r} contains a reduced expression for ww if and only if δ(𝐬𝐫)w\delta(\mathbf{s}\setminus\mathbf{r})\geq w [KM04, Lemma 3.4 (1)]. The dimension of face 𝐫\mathbf{r} is equal to (𝐫)1\ell(\mathbf{r})-1. Thus {𝐫𝐬,δ(𝐬𝐫)w}(1)(𝐫)1=χ~(Δ(𝐬,w))\sum_{\{\mathbf{r}\subseteq\mathbf{s},\delta(\mathbf{s}\setminus\mathbf{r})\geq w\}}(-1)^{\ell(\mathbf{r})-1}=\widetilde{\chi}(\Delta(\mathbf{s},w)). By the second statement of Theorem 4.3, {𝐫𝐬,δ(𝐬𝐫)>w}(1)(𝐫)1=χ~(Δ(𝐬,w))\sum_{\{\mathbf{r}\subseteq\mathbf{s},\delta(\mathbf{s}\setminus\mathbf{r})>w\}}(-1)^{\ell(\mathbf{r})-1}=\widetilde{\chi}(\partial\Delta(\mathbf{s},w)). Hence

𝐫𝐬δ(𝐬𝐫)=w(1)(𝐫)1=χ~(Δ(𝐬,w)).\sum_{\mathbf{r}\subseteq\mathbf{s}\atop\delta(\mathbf{s}\setminus\mathbf{r})=w}(-1)^{\ell(\mathbf{r})-1}=\widetilde{\chi}^{\circ}(\Delta(\mathbf{s},w)).

But since Δ(𝐬,w)\Delta(\mathbf{s},w) is either a ball or sphere, χ~(Δ(𝐬,w))=(1)dimΔ(𝐬,w)\widetilde{\chi}^{\circ}(\Delta(\mathbf{s},w))=(-1)^{\dim\Delta(\mathbf{s},w)}, by Lemma 4.1. To compute dimΔ(𝐬,w)\dim\Delta(\mathbf{s},w), observe that if 𝐫Δ(𝐬,w)\mathbf{r}\in\Delta(\mathbf{s},w) has maximal length, then 𝐬𝐫\mathbf{s}\setminus\mathbf{r} is a reduced word for ww; thus (𝐬)(𝐫)=(w)\ell(\mathbf{s})-\ell(\mathbf{r})=\ell(w), so (𝐫)=(𝐬)(w)\ell(\mathbf{r})=\ell(\mathbf{s})-\ell(w). Thus dimΔ(𝐬,w)=(𝐬)(w)1\dim\Delta(\mathbf{s},w)=\ell(\mathbf{s})-\ell(w)-1. We conclude

𝐫𝐬δ(𝐬𝐫)=w(1)l(𝐫)1=(1)(𝐬)(w)1.\sum_{\mathbf{r}\subseteq\mathbf{s}\atop\delta(\mathbf{s}\setminus\mathbf{r})=w}(-1)^{l(\mathbf{r})-1}=(-1)^{\ell(\mathbf{s})-\ell(w)-1}.

Multiplying both sides of this equation by (1)(𝐬)(w)1(-1)^{\ell(\mathbf{s})-\ell(w)-1}, we obtain

𝐫𝐬δ(𝐬𝐫)=w(1)e(𝐬𝐫)=1,\sum_{\mathbf{r}\subseteq\mathbf{s}\atop\delta(\mathbf{s}\setminus\mathbf{r})=w}(-1)^{e(\mathbf{s}\setminus\mathbf{r})}=1,

since, if δ(𝐬𝐫)=w\delta(\mathbf{s}\setminus\mathbf{r})=w, then we have (1)(𝐬)(w)1(1)(𝐫)1=(1)(𝐬)(𝐫)(w)=(1)(𝐬𝐫)(δ(𝐬𝐫))=(1)e(𝐬𝐫)(-1)^{\ell(\mathbf{s})-\ell(w)-1}(-1)^{\ell(\mathbf{r})-1}=(-1)^{\ell(\mathbf{s})-\ell(\mathbf{r})-\ell(w)}=(-1)^{\ell(\mathbf{s}\setminus\mathbf{r})-\ell(\delta(\mathbf{s}\setminus\mathbf{r}))}=(-1)^{e(\mathbf{s}\setminus\mathbf{r})}. Now the desired equation is obtained by re-indexing this summation. Rather than summing over subsequences 𝐫\mathbf{r} of 𝐬\mathbf{s}, one sums over their complementary subsequences 𝐪\mathbf{q}. ∎

Definition 4.5.

Let wWw\in W and let 𝐬=(s1,,sp)\mathbf{s}=(s_{1},\ldots,s_{p}) be a sequence of simple reflections in SS. Define 𝒯w,𝐬\mathcal{T}_{w,\mathbf{s}} to be the set of sequences 𝐭=(i1,,im)\mathbf{t}=(i_{1},\ldots,i_{m}), 1i1<<imp1\leq i_{1}<\cdots<i_{m}\leq p, such that Hsi1Hsim=HwH_{s_{i_{1}}}\cdots H_{s_{i_{m}}}=H_{w}. Then (𝐭)=m\ell(\mathbf{t})=m and e(𝐭)=(𝐭)(w)e(\mathbf{t})=\ell(\mathbf{t})-\ell(w).

Corollary 4.6.

𝐭𝒯w,𝐬(1)e(𝐭)=1\sum_{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}}(-1)^{e(\mathbf{t})}=1, if 𝒯w,𝐬\mathcal{T}_{w,\mathbf{s}}\neq\emptyset.

Proof.

We have

𝐭𝒯w,𝐬(1)e(𝐭)=𝐪𝐬δ(𝐪)=w(1)e(𝐪)=1,\sum_{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}}(-1)^{e(\mathbf{t})}=\sum_{\mathbf{q}\subseteq\mathbf{s}\atop\delta(\mathbf{q})=w}(-1)^{e(\mathbf{q})}=1,

where the first equality is obtained by re-indexing and the second equality is Corollary 4.4. Note that the hypothesis Tw,𝐬T_{w,\mathbf{s}}\neq\emptyset assures us that 𝐬\mathbf{s} contains a reduced expression for ww, a requirement for Corollary 4.4. ∎

We remark that the expression 𝐭𝒯w,𝐬(1)e(𝐭)\sum_{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}}(-1)^{e(\mathbf{t})} appearing in Corollary 4.6 has elements in common with the expression for ix[𝒪Xw]G/Bi_{x}^{*}[\mathcal{O}_{X^{w}}]_{G/B} given by (5.1). The similarity between these expressions is critical to our proof of Theorem 5.6.

5. Applications to Kazhdan-Lusztig and Schubert varieties

In Section 3 we saw that the pullback iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V} can be used to determine whether an integrally indecomposable weight αΦamb\alpha\in\Phi_{\operatorname{amb}} lies in Φnor\Phi_{\operatorname{nor}} or Φtan\Phi_{\tan}. Specifically, α\alpha lies in Φnor\Phi_{\operatorname{nor}} if and only if 1eθ1-e^{-\theta} is a simple factor of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V}. Computationally, however, an algorithm which utilizes this idea would seem to present difficulties, since it is often possible to express iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V} in many different ways. Determining whether 1eθ1-e^{-\theta} is a factor of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V} is nontrivial in general.

In this section we show that when YY is the Kazhdan-Lusztig variety YxwY_{x}^{w} in an appropriate space VG/BV\subseteq G/B, then a particular expression Pw,𝐬P_{w,\mathbf{s}} for iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V} due to Graham and Willems has the property that, if we assume that θ\theta is integrally indecomposable in Φamb=Φ(V)\Phi_{\operatorname{amb}}=\Phi(V), then whenever 1eθ1-e^{-\theta} is a simple factor of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V}, it is a factor of Pw,𝐬P_{w,\mathbf{s}} in a trivial fashion (see Theorem 5.6). Thus the expression Pw,𝐬P_{w,\mathbf{s}} allows us to detect simple factors 1eθ1-e^{-\theta} of iV[𝒪Y]Vi_{V}^{*}[\mathcal{O}_{Y}]_{V}, θ\theta integrally indecomposable, in a computationally simple manner.

We begin with two subsections reviewing properties of Kazhdan-Lusztig and Schubert varieties in G/BG/B.

5.1. Unipotent subgroups and affine spaces in G/BG/B

Let GG be a semisimple algebraic group defined over a algebraically closed field kk of characteristic 0, BB a Borel subgroup, BB^{-} the opposite Borel subgroup, and T=BBT=B\cap B^{-}, a maximal torus. Let W=NG(T)/TW=N_{G}(T)/T, the Weyl group of GG, and let SS be the set of simple reflections of WW relative to BB.

We consider several unipotent subgroups of GG, referring the reader to [Bor91, 14.12] for a more detailed discussion of their properties. Let UU and UU^{-} be the unipotent radicals of BB and BB^{-} respectively. Their weights, Φ(U)\Phi(U) and Φ(U)\Phi(U^{-}), are by definition the positive and negative roots Φ+\Phi^{+} and Φ\Phi^{-} respectively. The unipotent subgroup xUx1xU^{-}x^{-1}, which we denote by U(x)U^{-}(x), has weights xΦx\Phi^{-}. The unipotent subgroup U(x)UU^{-}(x)\cap U has weights Φ(U(x)U)=xΦΦ+\Phi(U^{-}(x)\cap U)=x\Phi^{-}\cap\Phi^{+}, which equals {αΦ+x1(α)<0}\{\alpha\in\Phi^{+}\mid x^{-1}(\alpha)<0\}, the inversion set of x1x^{-1}. Similarly, U(x)UU^{-}(x)\cap U^{-} has weights Φ(U(x)U)=xΦΦ={αΦx1(α)<0}={αΦ+x1(α)>0}=(Φ+I(x1))\Phi(U^{-}(x)\cap U^{-})=x\Phi^{-}\cap\Phi^{-}=\{\alpha\in\Phi^{-}\mid x^{-1}(\alpha)<0\}=-\{\alpha\in\Phi^{+}\mid x^{-1}(\alpha)>0\}=-(\Phi^{+}\setminus I(x^{-1})). The unipotent subgroups discussed in this paragraph are isomorphic to their Lie algebras [Bor91, Remark 14.4]. In particular, they are isomorphic to affine spaces, with which we often associate them.

The variety G/BG/B is called the full flag variety. The TT-fixed points of G/BG/B are of the form uBuB, uWu\in W. By Bruhat decomposition, under the mapping ζ:U(x)G/B\zeta:U^{-}(x)\to G/B, yyxBy\mapsto y\cdot xB, the unipotent subgroup U(x)U^{-}(x) embeds as a TT-stable affine space in G/BG/B containing xBxB. We denote this affine space by CxC_{x}. The unipotent subgroup U(x)UU^{-}(x)\cap U embeds as an affine subspace, which we denote by VV. We note that the weight spaces of CxC_{x} are one dimensional and that the weights lie in an open half-space; thus the same is true of VV.

5.2. Schubert and Kazhdan-Lusztig varieties in G/BG/B

The Schubert variety XwG/BX^{w}\subseteq G/B is defined to be BwB¯\overline{B^{-}wB}, the Zariski closure of the BB^{-} orbit through wBwB. The Kazhdan-Lusztig variety YxwY_{x}^{w} is defined to be VXwV\cap X^{w}. As the following lemma shows, locally, the two varieties differ only by an affine space with well-prescribed weights.

Lemma 5.1.

Let wxWw\leq x\in W. Then

  1. (i)

    Cx(U(x)U)×VC_{x}\cong(U^{-}(x)\cap U^{-})\times V.

  2. (ii)

    XwCx(U(x)U)×YxwX^{w}\cap C_{x}\cong(U^{-}(x)\cap U^{-})\times Y_{x}^{w}.

  3. (iii)

    TxXw(U(x)U)×TxYxwT_{x}X^{w}\cong(U^{-}(x)\cap U^{-})\times T_{x}Y_{x}^{w}.

  4. (iv)

    Φ(TxXw)=(Φ+I(x1))Φ(TxYxw)\Phi(T_{x}X^{w})=-(\Phi^{+}\setminus I(x^{-1}))\sqcup\Phi(T_{x}Y_{x}^{w}).

The isomorphisms (i) - (iii) are TT-equivariant isomorphisms of varieties.

Proof.

(i) [GK17, (4.6)].

(ii) is an application of [GK17, Lemma 4.6], with H=(U(x)U)H=(U^{-}(x)\cap U^{-}), Y=(U(x)U)xBY=(U^{-}(x)\cap U)\cdot xB, and Z=XwCxZ=X^{w}\cap C_{x}.

(iii) follows from (ii) and the fact that the tangent space of a product is isomorphic to the product of the tangent spaces (see [GW10, Proposition 6.9]).

(iv) We saw in Sectionn 5.1 that Φ(U(x)U)=(Φ+I(x1))\Phi(U^{-}(x)\cap U^{-})=-(\Phi^{+}\setminus I(x^{-1})). Combined with (iii), this yields the desired result. ∎

Lemma 5.1(iv) shows how to produce Φ(TxXw)\Phi(T_{x}X^{w}) from Φ(TxYxw)\Phi(T_{x}Y_{x}^{w}). In Section 3, we saw that information about the latter can be obtained from iV[𝒪Yxw]Vi_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}. The next proposition asserts that this pullback is equal to ix[𝒪Xw]G/Bi_{x}^{*}[\mathcal{O}_{X^{w}}]_{G/B}, for which there are known formulas, in particular (5.1) below.

Lemma 5.2.

Let wxWw\leq x\in W. Then iV[𝒪Yxw]V=ix[𝒪Xw]G/Bi_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}=i_{x}^{*}[\mathcal{O}_{X^{w}}]_{G/B}.

Proof.

By Lemma 5.1(i) and (ii), we can apply [GK17, Lemma 2.1] to obtain iV[𝒪Yxw]V=iCx[𝒪XwCx]Cxi_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}=i_{C_{x}}^{*}[\mathcal{O}_{X^{w}\cap C_{x}}]_{C_{x}}. Since pullbacks in equivariant KK-theory are defined locally, iCx[𝒪XwCx]Cx=ix[𝒪Xw]G/Bi_{C_{x}}^{*}[\mathcal{O}_{X^{w}\cap C_{x}}]_{C_{x}}=i_{x}^{*}[\mathcal{O}_{X^{w}}]_{G/B}. ∎

5.3. Formulas for weights of the tangent space

Fix a reduced expression 𝐬=(s1,,sl)\mathbf{s}=(s_{1},\ldots,s_{l}) for xx. The elements of the inversion set I(x1)=Φ+xΦI(x^{-1})=\Phi^{+}\cap x\Phi^{-} are given explicitly by the formula γi=s1si(αi)\gamma_{i}=s_{1}\cdots s_{i}(\alpha_{i}), i=1,,li=1,\ldots,l, where αi\alpha_{i} is the simple root corresponding to sis_{i} [Hum90]. The following result is due to Graham [Gra02] and Willems [Wil06]:

Theorem 5.3.

Let wxWw\leq x\in W, and let 𝐬=(s1,,sl)\mathbf{s}=(s_{1},\ldots,s_{l}) be a reduced sequence of simple reflections for xx. Then

ix[𝒪Xw]G/B=𝐭𝒯w,𝐬(1)e(𝐭)i𝐭(1eγi)R(T)i_{x}^{*}[\mathcal{O}_{X^{w}}]_{G/B}=\sum_{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}}(-1)^{e(\mathbf{t})}\prod_{i\in\mathbf{t}}(1-e^{-\gamma_{i}})\in R(T) (5.1)

Denote the expression 𝐭𝒯w,𝐬(1)e(𝐭)i𝐭(1eγi)\sum_{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}}(-1)^{e(\mathbf{t})}\prod_{i\in\mathbf{t}}(1-e^{-\gamma_{i}}) by Pw,𝐬P_{w,\mathbf{s}}. By Lemma 5.2, we have

Corollary 5.4.

Let wxWw\leq x\in W. Then iV[𝒪Yxw]V=Pw,𝐬i_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}=P_{w,\mathbf{s}}.

Remark 5.5.

In general, there exist numerous expressions for iV[𝒪Yxw]Vi_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}. Lemma 2.2 assures us that there exists an expression as a polynomial in 1eγ1-e^{-\gamma}, γΦ(V)=I(x1)\gamma\in\Phi(V)=I(x^{-1}); Pw,𝐬P_{w,\mathbf{s}} is such an expression.

We shall say that 1eγj1-e^{-\gamma_{j}} is an explicit factor of Pw,𝐬P_{w,\mathbf{s}} if 1eγj1-e^{-\gamma_{j}} occurs among the factors of every summand i𝐭(1eγi)\prod_{i\in\mathbf{t}}(1-e^{-\gamma_{i}}) of Pw,𝐬P_{w,\mathbf{s}}, or equivalently, if jj belongs to every 𝐭𝒯w,𝐬\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}} (see Remark 3.2). Since all of the γj\gamma_{j}, j=1,,lj=1,\ldots,l, are distinct, every explicit factor of Pw,𝐬P_{w,\mathbf{s}} is a simple factor of iV[𝒪Yxw]Vi_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}. The following theorem tells us that when γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}), the converse is true as well.

Theorem 5.6.

Let wxWw\leq x\in W, and let γj\gamma_{j} be integrally indecomposable in I(x1)I(x^{-1}). If 1eγj1-e^{-\gamma_{j}} is a simple factor of iV[𝒪Yxw]Vi_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}, then it is an explicit factor of Pw,𝐬P_{w,\mathbf{s}}.

Proof.

Let CC be the coordinate ring of the tangent cone to YxwY_{x}^{w} at xx. We will assume that 1eγj1-e^{-\gamma_{j}} is not an explicit factor of Pw,𝐬P_{w,\mathbf{s}} and show that γj-\gamma_{j} is a weight of CC (of multiplicity 1). Proposition 3.5 then implies that 1eγj1-e^{-\gamma_{j}} is not a simple factor of iV[𝒪Yxw]Vi_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}, completing the proof. Let [l][l] denote {1,,l}\{1,\ldots,l\}.

By Proposition 2.3, we have

CharC=iV[𝒪Yxw]Vλ1(V)=𝐭𝒯w,𝐬(1)e(𝐭)i𝐭(1eγi)i[l](1eγi)\operatorname{Char}C=\frac{i_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}}{\lambda_{-1}(V^{*})}=\frac{\sum_{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}}(-1)^{e(\mathbf{t})}\prod_{i\in\mathbf{t}}(1-e^{-\gamma_{i}})}{\prod_{i\in[l]}(1-e^{-\gamma_{i}})} (5.2)

Each summand of (5.2) can be simplified:

(1)e(𝐭)i𝐭(1eγi)i[l](1eγi)\displaystyle\frac{(-1)^{e(\mathbf{t})}\prod_{i\in\mathbf{t}}(1-e^{-\gamma_{i}})}{\prod_{i\in[l]}(1-e^{-\gamma_{i}})} =(1)e(𝐭)1i𝐭(1eγi)\displaystyle=(-1)^{e(\mathbf{t})}\frac{1}{\prod_{i\notin\mathbf{t}}(1-e^{-\gamma_{i}})}
=(1)e(𝐭)i𝐭(1+eγi+e2γi+)\displaystyle=(-1)^{e(\mathbf{t})}\prod_{i\notin\mathbf{t}}(1+e^{-\gamma_{i}}+e^{-2\gamma_{i}}+\cdots)
=(1)e(𝐭)ζCone{γi:i𝐭}nζeζ\displaystyle=(-1)^{e(\mathbf{t})}\sum\nolimits_{\zeta\in\operatorname{Cone}_{\mathbb{Z}}\{\gamma_{i}:i\notin\mathbf{t}\}}n_{\zeta}e^{-\zeta}
=(1)e(𝐭)nγjeγj+ other terms\displaystyle=(-1)^{e(\mathbf{t})}n_{\gamma_{j}}e^{-\gamma_{j}}+\text{ other terms}

where nζn_{\zeta} is the number of ways to express ζ\zeta as a nonnegative integer linear combination of the γi\gamma_{i}, i𝐭i\notin\mathbf{t}, and “other terms” refers to an infinite linear combination of characters with no eγje^{-\gamma_{j}} term. Since γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}), nγj=1n_{\gamma_{j}}=1 if j𝐭j\notin\mathbf{t} and nγj=0n_{\gamma_{j}}=0 if j𝐭j\in\mathbf{t}. Thus

(1)e(𝐭)i𝐭(1eγi)i[l](1eγi)={(1)e(𝐭)eγj+ other terms,if j𝐭other terms,if j𝐭\frac{(-1)^{e(\mathbf{t})}\prod_{i\in\mathbf{t}}(1-e^{-\gamma_{i}})}{\prod_{i\in[l]}(1-e^{-\gamma_{i}})}=\begin{cases}(-1)^{e(\mathbf{t})}e^{-\gamma_{j}}+\text{ other terms},&\text{if }j\notin\mathbf{t}\\ \text{other terms},&\text{if }j\in\mathbf{t}\end{cases} (5.3)

According to (5.2), CharC\operatorname{Char}C is the sum of fractions as in (5.3), one for each 𝐭𝒯w,𝐬\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}. Therefore the coefficient of eγje^{-\gamma_{j}} in CharC\operatorname{Char}C is

N={𝐭𝒯w,𝐬:j𝐭}(1)e(𝐭)N=\sum_{\{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}}:j\notin\mathbf{t}\}}(-1)^{e(\mathbf{t})}

Setting 𝐬j=(s1,,s^j,,sl)\mathbf{s}_{j}=(s_{1},\ldots,\hat{s}_{j},\ldots,s_{l}), we have

N=𝐭𝒯w,𝐬j(1)e(𝐭)N=\sum_{\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}_{j}}}(-1)^{e(\mathbf{t})}

The assumption that 1eγj1-e^{-\gamma_{j}} is not an explicit factor of Pw,𝐬P_{w,\mathbf{s}} assures us that 𝒯w,𝐬j\mathcal{T}_{w,\mathbf{s}_{j}}\neq\emptyset, and thus this sum equals 1 by Corollary 4.6. Since N0N\neq 0, γj-\gamma_{j} is a weight of CC. ∎

Denote Φ(TxYxw)\Phi(T_{x}Y_{x}^{w}) and Φ(V/TxYxw)\Phi(V/T_{x}Y_{x}^{w}) by Φtan\Phi_{\tan} and Φnor\Phi_{\operatorname{nor}} respectively, so ΦtanΦnor=Φ(V)=I(x1)\Phi_{\tan}\sqcup\Phi_{\operatorname{nor}}=\Phi(V)=I(x^{-1}).

Corollary 5.7.

Let wxWw\leq x\in W, and suppose that γ\gamma is an integrally indecomposable element of I(x1)I(x^{-1}). If γΦnor\gamma\in\Phi_{\operatorname{nor}}, then 1eγ1-e^{-\gamma} is an explicit factor of Pw,𝐬P_{w,\mathbf{s}}.

Proof.

Since γΦnor\gamma\in\Phi_{\operatorname{nor}}, 1eγ1-e^{-\gamma} is a simple factor of iV[𝒪Yxw]Vi_{V}^{*}[\mathcal{O}_{Y_{x}^{w}}]_{V}, by Proposition 3.3. Thus 1eγ1-e^{-\gamma} is an explicit factor of Pw,𝐬P_{w,\mathbf{s}}, by Theorem 5.6. ∎

Let m=(w)m=\ell(w), and define

𝒯w,𝐬={𝐭=(t1,,tm)[l]st1stm=w}.\mathcal{R}\mathcal{T}_{w,\mathbf{s}}=\{\mathbf{t}=(t_{1},\ldots,t_{m})\subseteq[l]\mid s_{t_{1}}\cdots s_{t_{m}}=w\}.

Parts (i) - (iii) of the following theorem summarize the main findings of this section thus far. Parts (iv) and (v) provide a computationally simpler method of determining whether γj\gamma_{j} lies in Φtan\Phi_{\tan}, by allowing us to substitute 𝒯w,𝐬\mathcal{R}\mathcal{T}_{w,\mathbf{s}} for 𝒯w,𝐬\mathcal{T}_{w,\mathbf{s}}, and thus to perform calculations in the Weyl group rather than the 0-Hecke algebra. Part (vi) gives an alternative characterization of (v) in terms of Demazure products.

Theorem 5.8.

Let wxWw\leq x\in W, and let 𝐬=(s1,,sl)\mathbf{s}=(s_{1},\ldots,s_{l}) be a reduced expression for xx. If γj\gamma_{j} is an integrally indecomposable element of I(x1)I(x^{-1}), then the following are equivalent:

  1. (i)

    γjΦtan\gamma_{j}\in\Phi_{\tan}.

  2. (ii)

    1eγj1-e^{-\gamma_{j}} is not an explicit factor of Pw,𝐬P_{w,\mathbf{s}}.

  3. (iii)

    There exists 𝐭𝒯w,𝐬\mathbf{t}\in\mathcal{T}_{w,\mathbf{s}} not containing jj.

  4. (iv)

    There exists 𝐭𝒯w,𝐬\mathbf{t}\in\mathcal{R}\mathcal{T}_{w,\mathbf{s}} not containing jj.

  5. (v)

    There exists a reduced subexpression of (s1,,s^j,,sl)(s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}) for ww.

  6. (vi)

    δ((s1,,s^j,,sl))w\delta((s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}))\geq w.

Proof.

(i) \Leftrightarrow (ii) by Proposition 3.5 and Corollary 5.7; (ii) \Leftrightarrow (iii) and (iv) \Leftrightarrow (v) are due to definitions of Pw,𝐬P_{w,\mathbf{s}}, 𝒯w,𝐬\mathcal{T}_{w,\mathbf{s}}, and 𝒯w,𝐬\mathcal{R}\mathcal{T}_{w,\mathbf{s}}. The proof of (iii) \Leftrightarrow (iv) follows from 𝒯w,𝐬𝒯w,𝐬\mathcal{T}_{w,\mathbf{s}}\supseteq\mathcal{R}\mathcal{T}_{w,\mathbf{s}} and the fact that every element of 𝒯w,𝐬\mathcal{T}_{w,\mathbf{s}} contains an element of 𝒯w,𝐬\mathcal{R}\mathcal{T}_{w,\mathbf{s}}.

(v) \Leftrightarrow (vi) There exists a reduced subexpression of 𝐬\mathbf{s} for ww not containing sjs_{j} if and only if there exists a subexpression of (s1,,s^j,,sl)(s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}) for ww if and only if δ((s1,,s^j,,sl))w\delta((s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}))\geq w, where the last equivalence is due to [KM04, Lemma 3.4 (1)]. ∎

Remark 5.9.

For γjI(x1)\gamma_{j}\in I(x^{-1}), it is known that in type AA, γjΦtan\gamma_{j}\in\Phi_{\tan} if and only if s1s^jslws_{1}\cdots\widehat{s}_{j}\cdots s_{l}\geq w [LS84]. Theorem 5.8 states that if γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}), then γjΦtan\gamma_{j}\in\Phi_{\tan} if and only if δ((s1,,s^j,,sl))w\delta((s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}))\geq w. These two statements imply that in type AA, if γj\gamma_{j} is integrally indecomposable in I(x1)I(x^{-1}), then δ((s1,,s^j,,sl))w\delta((s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}))\geq w if and only if s1s^jslws_{1}\cdots\widehat{s}_{j}\cdots s_{l}\geq w. That this holds for all wxw\leq x would seem to imply that δ((s1,,s^j,,sl))=s1s^jsl\delta((s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}))=s_{1}\cdots\widehat{s}_{j}\cdots s_{l}. This is indeed true, and the above argument can be made rigorous. In [GK21], it is shown that the statement extends to all simply-laced types. It is also shown that if γj\gamma_{j} is rationally indecomposable in I(x1)I(x^{-1}), then δ((s1,,s^j,,sl))=s1s^jsl\delta((s_{1},\ldots,\widehat{s}_{j},\ldots,s_{l}))=s_{1}\cdots\widehat{s}_{j}\cdots s_{l} in all types.

Remark 5.10.

Suppose that γj\gamma_{j} is not integrally indecomposable in I(x1)I(x^{-1}). Then statements (ii) – (vi) of Theorem 5.8 are still equivalent, but statement (i) is no longer equivalent to the other five in general. The following example shows that (vi) \Rightarrow (i) can fail. In type A2A_{2}, let 𝐬=(σ1,σ2,σ1)\mathbf{s}=(\sigma_{1},\sigma_{2},\sigma_{1}), where σi\sigma_{i} is the simple transposition which exchanges ii and i+1i+1. Let w=σ1w=\sigma_{1} and j=2j=2. Then δ((σ1,σ^2,σ1))=δ((σ1,σ1))=σ1w\delta((\sigma_{1},\widehat{\sigma}_{2},\sigma_{1}))=\delta((\sigma_{1},\sigma_{1}))=\sigma_{1}\geq w, so (vi) holds. However, σ1σ^2σ1=ew\sigma_{1}\widehat{\sigma}_{2}\sigma_{1}=e\not\geq w. Thus γ2Φtan\gamma_{2}\notin\Phi_{\tan} (see Remark 5.9), so (i) fails.

We note that γj\gamma_{j} is required to be integrally indecomposable in I(x1)I(x^{-1}) for our proofs of both implications of Theorem 5.8 (i) \Leftrightarrow (ii).

6. Partial flag varieties and cominuscule elements

Let PP be a parabolic subgroup containing BB. In Section 6.1 we show that Lemma 5.1 and Theorem 5.8 extend from G/BG/B to G/PG/P with no changes other than notation. In Section 6.2 we apply the results to cominuscule elements of WW and cominuscule G/PG/P.

6.1. Extending results to G/PG/P

Let PP be a parabolic subgroup containing BB. Let LL be the Levi subgroup of PP containing TT, and WP=NL(T)/TW_{P}=N_{L}(T)/T, the Weyl group of LL. Each coset uWPuW_{P} in W/WPW/W_{P} contains a unique representative of minimal length; denote the set of minimal length coset representatives by WPWW^{P}\subseteq W. Unless stated otherwise, in this subsection we assume that all Weyl group elements lie in WPW^{P}. The TT-fixed points of G/PG/P are of the form uPuP, uWPu\in W^{P}.

Let PP^{-} the opposite parabolic subgroup to PP, and let UPU_{P}^{-} be the the unipotent radical of PP^{-}. Under the mapping ζ:UP(x)G/P\zeta:U_{P}^{-}(x)\to G/P, yyxPy\mapsto y\cdot xP, the unipotent subgroup UP(x)U_{P}^{-}(x) embeds as a TT-stable affine space in G/PG/P containing xPxP. The unipotent subgroup UP(x)UU_{P}^{-}(x)\cap U embeds as an affine subspace, which we denote by VPV_{P}.

The Schubert variety XPwG/PX_{P}^{w}\subseteq G/P is defined to be BwP¯\overline{B^{-}wP}, the Zariski closure of the BB^{-} orbit through wPwP. The Kazhdan-Lusztig variety Yx,PwY_{x,P}^{w} is defined to be VPXPwV_{P}\cap X_{P}^{w}.

The following result appears in [Knu09, Section 7.3]:

Theorem 6.1.

Let wxWPw\leq x\in W^{P}. Then VPVV_{P}\cong V and Yx,PwYxwY_{x,P}^{w}\cong Y_{x}^{w}.

The next theorem extends the main results of Section 5 to G/PG/P.

Theorem 6.2.

Let wxWPw\leq x\in W^{P}.

  1. (i)

    Φ(TxXPw)=Φ(UP(x)U)Φ(TxYx,Pw)\Phi(T_{x}X_{P}^{w})=\Phi(U_{P}^{-}(x)\cap U^{-})\sqcup\Phi(T_{x}Y_{x,P}^{w}).

  2. (ii)

    Let γj\gamma_{j} be an integrally indecomposable element of I(x1)I(x^{-1}). Then γjΦ(TxYx,Pw)\gamma_{j}\in\Phi(T_{x}Y_{x,P}^{w}) if and only if δ((s1,,s^j,,sl))w\delta((s_{1},\ldots,\hat{s}_{j},\ldots,s_{l}))\geq w.

Proof.

(i) Lemma 5.1(i)-(iii) remain valid if all quantities are replaced by their analogs in G/PG/P. In particular, TxXPw(UP(x)U)×TxYx,PwT_{x}X_{P}^{w}\cong(U_{P}^{-}(x)\cap U^{-})\times T_{x}Y_{x,P}^{w}.

(ii) Since Yx,PwYxwY_{x,P}^{w}\cong Y_{x}^{w}, TxYx,PwTxYxwT_{x}Y_{x,P}^{w}\cong T_{x}Y_{x}^{w}. Thus all parts of Theorem 5.8 remain valid if Φtan=Φ(TxYxw)\Phi_{\tan}=\Phi(T_{x}Y_{x}^{w}) in Theorem 5.8(i) is replaced by Φ(TxYx,Pw)\Phi(T_{x}Y_{x,P}^{w}). ∎

6.2. Application to cominuscule Weyl group elements and cominuscule G/PG/P

In this subsection we discuss conditions on xx under which all elements of I(x1)I(x^{-1}) are integrally indecomposable, and thus, for any Kazhdan-Lusztig variety containing xx, Theorems 5.8 and 6.2(ii) recover all weights of the tangent space at xx. In particular, we show that our results completely describe the tangent spaces of Schubert varieties in cominuscule G/PG/P.

Definition 6.3.

The element xWx\in W is said to be cominuscule if there exists v𝔱v\in{\mathfrak{t}} such that α(v)=1\alpha(v)=-1 for all αI(x1)\alpha\in I(x^{-1}).

This notion was introduced and studied by Peterson (see [GK17, Section 5.2] or [Ste01] for discussion). In type AA, the cominuscule Weyl group elements are precisely the 321-avoiding permutations [Knu09, p. 25]. As noted in [GK17], the equality I(x)=x1I(x1)I(x)=-x^{-1}I(x^{-1}) implies that xx is cominuscule if and only if x1x^{-1} is.

Proposition 6.4.

If xWx\in W is cominuscule, then all elements of I(x1)I(x^{-1}) are integrally indecomposable.

Proof.

If xx is cominuscule, then there exists v𝔱v\in{\mathfrak{t}} such that α(v)=1\alpha(v)=-1 for all αI(x1)\alpha\in I(x^{-1}). Assume that some βI(x1)\beta\in I(x^{-1}) is integrally decomposable. Then β=i=1mβi\beta=\sum_{i=1}^{m}\beta_{i}, where m2m\geq 2, βiI(x1)\beta_{i}\in I(x^{-1}). Since β(v)=1\beta(v)=-1 and βi(v)=1\beta_{i}(v)=-1 for all ii, this leads to a contradiction. ∎

Remark 6.5.

The converse of the above proposition is false: there exist non-cominuscule elements xx such that every element of I(x1)I(x^{-1}) is integrally indecomposable. The following example is a variation and extension of [Ste01, Remark 5.4]. In type D4D_{4}, with the conventions of [Hum90], consider the element x=s2s1s3s4s2x=s_{2}s_{1}s_{3}s_{4}s_{2}. The inversion set I(x1)I(x^{-1}) is equal to {ϵ1ϵ3,ϵ1+ϵ2,ϵ2ϵ3,ϵ2ϵ4,ϵ2+ϵ4}\{\epsilon_{1}-\epsilon_{3},\epsilon_{1}+\epsilon_{2},\epsilon_{2}-\epsilon_{3},\epsilon_{2}-\epsilon_{4},\epsilon_{2}+\epsilon_{4}\}. Every element of I(x1)I(x^{-1}) is integrally indecomposable, but the element xx is not cominuscule (cf. [Ste01, Remark 5.4]). Note that [Ste01] uses a different numbering of the vertices of the Dynkin diagram in which node 33 has degree 33 (see [Ste01, Remark 2.7]), so he writes the element xx as s3s1s2s4s3s_{3}s_{1}s_{2}s_{4}s_{3}.

Definition 6.6.

The maximal parabolic subgroup PBP\supseteq B is said to be cominuscule if the simple root αi\alpha_{i} corresponding to PP occurs with coefficient 1 when the highest root of GG is written as a linear combination of the simple roots.

If PP is cominuscule, then the corresponding flag variety G/PG/P is said to be cominuscule as well. We refer the reader to [BL00, Chapter 9], [Bou02, Chapter VI, §1, Exercise 24], [GK15] for more on cominuscule G/PG/P. The following proposition gives an important class of cominuscule Weyl group elements.

Proposition 6.7.

If xWPx\in W^{P}, where PP is cominuscule, then xx is a cominuscule element of WW.

Proof.

If xWPx\in W^{P}, then U(x)U=UP(x)UU^{-}(x)\cap U=U_{P}^{-}(x)\cap U (see the discussion before Lemma 4.1 in [GK15], cf. [Knu09]). Hence

I(x1)=Φ((U(x)U)=Φ(UP(x)U)xΦ(UP).I(x^{-1})=\Phi((U^{-}(x)\cap U)=\Phi(U_{P}^{-}(x)\cap U)\subset x\Phi(U_{P}^{-}).

Let α1,,αr\alpha_{1},\ldots,\alpha_{r} denote the simple roots of GG; these form a basis for 𝔱{\mathfrak{t}}^{*}. Denote the dual basis of 𝔱{\mathfrak{t}} by ξ1,,ξr\xi_{1},\ldots,\xi_{r}. Assume that PP corresponds to the simple root αi\alpha_{i}. Since PP is cominuscule, [GK15, Lemma 2.8] implies αi\alpha_{i} must occur with coefficient 1-1 in all αΦ(UP)\alpha\in\Phi(U_{P}^{-}) (when α\alpha is written as a linear combination of the simple roots), so for all such α\alpha, we have α(ξi)=1\alpha(\xi_{i})=-1. It follows that v=xξiv=x\xi_{i} satisfies α(v)=1\alpha(v)=-1 for all αxΦ(UP)\alpha\in x\Phi(U_{P}^{-}). Hence α(v)=1\alpha(v)=-1 for all αI(x1)\alpha\in I(x^{-1}), so xx is a cominuscule element of WW. ∎

Remark 6.8.

The results of this subsection imply that if PP is cominuscule and wxWPw\leq x\in W^{P}, then Theorem 6.2 characterizes all weights of TxYx,PwT_{x}Y_{x,P}^{w} and TxXPwT_{x}X_{P}^{w}. More generally, suppose xWx\in W is any cominuscule element (or more generally any element such that each element of I(x1)I(x^{-1}) is integrally indecomposable). Then Lemma 5.1 and Theorem 5.8 characterize all weights of TxYxwT_{x}Y_{x}^{w} and TxXwT_{x}X^{w}. If in addition PBP\supset B is a parabolic subgroup such that w,xWPw,x\in W^{P}, then Theorem 6.2 characterizes all weights of TxYx,PwT_{x}Y_{x,P}^{w} and TxXPwT_{x}X_{P}^{w}.

References

  • [BL00] Sara Billey and V. Lakshmibai, Singular loci of Schubert varieties, Progress in Mathematics, vol. 182, Birkhäuser Boston Inc., Boston, MA, 2000. MR 1782635 (2001j:14065)
  • [Bor91] Armand Borel, Linear algebraic groups, second ed., Graduate Texts in Mathematics, vol. 126, Springer-Verlag, New York, 1991. MR 1102012
  • [Bou02] Nicolas Bourbaki, Lie groups and Lie algebras. Chapters 4–6, Elements of Mathematics (Berlin), Springer-Verlag, Berlin, 2002, Translated from the 1968 French original by Andrew Pressley. MR 1890629
  • [Car95] James B. Carrell, On the smooth points of a Schubert variety, Representations of groups (Banff, AB, 1994), CMS Conf. Proc., vol. 16, Amer. Math. Soc., Providence, RI, 1995, pp. 15–33. MR 1357193
  • [CK03] James B. Carrell and Jochen Kuttler, Smooth points of TT-stable varieties in G/BG/B and the Peterson map, Invent. Math. 151 (2003), no. 2, 353–379. MR 1953262
  • [GK15] William Graham and Victor Kreiman, Excited Young diagrams, equivariant KK-theory, and Schubert varieties, Trans. Amer. Math. Soc. 367 (2015), no. 9, 6597–6645. MR 3356949
  • [GK17] by same author, Cominuscule points and Schubert varieties, 2017.
  • [GK21] by same author, Tangent spaces and TT-invariant curves of Schubert varieties, 2021.
  • [Gra02] William Graham, Equivariant K{K}-theory and Schubert varieties, preprint (2002).
  • [GW10] Ulrich Görtz and Torsten Wedhorn, Algebraic geometry I, Advanced Lectures in Mathematics, Vieweg + Teubner, Wiesbaden, 2010, Schemes with examples and exercises. MR 2675155
  • [Hum90] James E. Humphreys, Reflection groups and Coxeter groups, Cambridge Studies in Advanced Mathematics, vol. 29, Cambridge University Press, Cambridge, 1990. MR 1066460 (92h:20002)
  • [KM04] Allen Knutson and Ezra Miller, Subword complexes in Coxeter groups, Adv. Math. 184 (2004), no. 1, 161–176. MR 2047852 (2005c:20066)
  • [KM05] by same author, Gröbner geometry of Schubert polynomials, Ann. of Math. (2) 161 (2005), no. 3, 1245–1318. MR 2180402 (2006i:05177)
  • [Knu09] Allen Knutson, Frobenius splitting, point-counting, and degeneration, 2009.
  • [Lak95] V. Lakshmibai, Tangent spaces to Schubert varieties, Math. Res. Lett. 2 (1995), no. 4, 473–477. MR 1355708
  • [Lak00a] by same author, On tangent spaces to Schubert varieties, J. Algebra 230 (2000), no. 1, 222–244. MR 1774765
  • [Lak00b] by same author, On tangent spaces to Schubert varieties. II, J. Algebra 224 (2000), no. 2, 167–197. MR 1739576
  • [LS84] V. Lakshmibai and C. S. Seshadri, Singular locus of a Schubert variety, Bull. Amer. Math. Soc. (N.S.) 11 (1984), no. 2, 363–366. MR 752799
  • [Mau80] Charles Richard Francis Maunder, Algebraic topology, Cambridge University Press, Cambridge-New York, 1980. MR 694843
  • [Ros89] W. Rossmann, Equivariant multiplicities on complex varieties, Astérisque (1989), no. 173-174, 11, 313–330, Orbites unipotentes et représentations, III. MR 91g:32042
  • [Ste01] John R. Stembridge, Minuscule elements of Weyl groups, J. Algebra 235 (2001), no. 2, 722–743. MR 1805477
  • [Wil06] Matthieu Willems, KK-théorie équivariante des tours de Bott. Application à la structure multiplicative de la KK-théorie équivariante des variétés de drapeaux, Duke Math. J. 132 (2006), no. 2, 271–309. MR 2219259 (2007b:19009)