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

From braids to transverse slices in reductive groups

Wicher Malten
Abstract

In 1965, Steinberg’s study of conjugacy classes in connected reductive groups led him to construct an affine subspace parametrising regular conjugacy classes, which he noticed is also a cross section for the conjugation action by the unipotent radical of a Borel subgroup on another affine subspace. Recently, generalisations of this slice and its cross section property have been obtained by Sevostyanov in the context of quantum group analogues of WW-algebras and by He-Lusztig in the context of Deligne-Lusztig varieties. Such slices are often thought of as group analogues of Slodowy slices.

In this paper we explain their relationship via common generalisations associated to Weyl group elements and provide a simple criterion for cross sections in terms of roots. In the most important class of examples this criterion is equivalent to a statement about the Deligne-Garside factors of their powers in the braid monoid being maximal in some sense. Moreover, we show that these subvarieties transversely intersect conjugacy classes and determine for a large class of factorisable rr-matrices when the Semenov-Tian-Shansky bracket reduces to a Poisson structure on these slices.

1 Introduction

In 1965, Steinberg’s study of conjugacy classes in a connected reductive group GG led him to analyse its regular elements, characterising them in various ways; they can be defined as the elements whose conjugacy class has minimal codimension (which is equal to the rank of GG), and together these elements form a dense open subset. Fixing a Borel subgroup BB and maximal torus TT, denoting by N+=[B,B]N_{+}=[B,B] the unipotent radical of BB, by NN_{-} its opposite, by W=NG(T)/TW=N_{G}(T)/T the Weyl group, writing for ww in WW then Nw:=N+w1NwN_{w}:=N_{+}\cap w^{-1}N_{-}w and w˙\dot{w} for arbitrary lifts to the normaliser NG(T)N_{G}(T) of TT in GG and finally fixing ww to be a Coxeter element of minimal length,111By Coxeter element we mean any conjugate of a product (in any order) of all the simple reflections. he proves that the Steinberg slice w˙Nw\dot{w}N_{w} yields a cross section of these regular conjugacy classes [Ste65, §1.4].

Example 1.1 ([Ste65, §7.4]).

Let G=SLrk+1G=\mathrm{SL}_{\mathrm{rk}+1} over a scheme SS and consider the Coxeter element w=srks1w=s_{\mathrm{rk}}\cdots s_{1}. The usual lift w˙\dot{w} yields the space of Frobenius companion matrices

w˙Nw={[010000100001(1)rkcrkc2c1]:c1,,crk𝒪S}.\dot{w}N_{w}=\left\{\begin{bmatrix}0&1&\cdots&0&0\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&\cdots&1&0\\ 0&0&\cdots&0&1\\ (-1)^{\mathrm{rk}}&c_{\mathrm{rk}}&\cdots&c_{2}&c_{1}\end{bmatrix}:c_{1},\ldots,c_{\mathrm{rk}}\in\mathcal{O}_{S}\right\}.

Furthermore, Steinberg settles the existence of regular unipotent elements by locating some of them inside of the Bruhat cell BwBBwB [Ste65, §4]. He subsequently remarks that actually BwBBwB consists entirely of regular elements, which he notes follows from the same property of w˙Nw\dot{w}N_{w} in combination with the isomorphism

N+×w˙NwN+w˙N+,(n,g)n1gn,N_{+}\times\dot{w}N_{w}\overset{\sim}{\longrightarrow}N_{+}\dot{w}N_{+},\qquad(n,g)\longmapsto n^{-1}gn, (1.1)

but the proof of that isomorphism is missing [Ste65, §8.9]. In the late 1970s this map was investigated by Spaltenstein, who found an example in type 𝖠5\mathsf{A}_{5} (see Example 1.15) showing that it is not necessarily an isomorphism when ww is replaced by a Coxeter element which is not of minimal length [HL12, §0.4] . Very recently, generalisations of this cross section appeared in two different settings:

Perhaps the most natural way to construct WW-algebras is by applying quantum Hamiltonian reduction to a certain character of the Lie algebra 𝔫\mathfrak{n} of a subgroup NN of N+N_{+} generated by root subgroups [dBT93, Pre02, GG02]. For example, Kostant’s study in 1978 of Whittaker representations for the Langlands program led him to construct the principal finite WW-algebra [Kos78], which can be interpreted as applying this procedure to a regular character of the Lie algebra 𝔫+\mathfrak{n}_{+} of N+N_{+}. The character of 𝔫\mathfrak{n} dequantises to a symplectic point of 𝔫\mathfrak{n}^{*} via Dixmier’s map [Dix63] and the semiclassical limit of the corresponding finite WW-algebra is then obtained by reducing the inverse image of this point under the momentum map for the restriction of the coadjoint action of GG on the dual 𝔤\mathfrak{g}^{*} of its Lie algebra to NN. The quantum group analogue of that reduces inverse images of symplectic points under momentum maps for the restriction of the dressing action of a factorisable Poisson-Lie group GG (with the Drinfeld-Sklyanin bracket) on its dual Poisson-Lie group GG^{*} (which is quantised by the Drinfeld-Jimbo quantum group UqGU_{q}G) to coisotropic subgroups NN of N+N_{+} generated by root subgroups. One can exponentiate the symplectic point in 𝔫+\mathfrak{n}_{+}^{*} to a point in the dual N+N_{+}^{*} of N+N_{+}, and show directly that it is symplectic if and only if the standard rr-matrix on GG is modified by the Cayley transform of a Coxeter element of minimal length [Mal]. Denoting by GG_{*} the variety GG equipped with the corresponding Semenov-Tian-Shansky bracket (which is quantised by the integrable part UqintGU_{q}^{\mathrm{int}}G of UqGU_{q}G), the resulting reduction in GG^{*} covers a reduction in GG_{*} along the GG-equivariant factorisation mapping GGG^{*}\rightarrow G_{*}. More precisely, this is the Poisson reduction of N+w˙N+GN_{+}\dot{w}N_{+}\subset G_{*} by the conjugation action of N+GN_{+}\subset G, so the cross section suggests that a certain reduction of UqintGU_{q}^{\mathrm{int}}G should quantise the Steinberg slice w˙Nw\dot{w}N_{w}. The differential graded algebra corresponding to the principal case was thoroughly studied in work of Sevostyanov [Sev00], the author [Mal], and Grojnowski [Gro]. Whilst searching for quantum group analogues of non-principal WW-algebras, Sevostyanov generalised the Steinberg slice to some elements in every conjugacy class of the Weyl group, and proved that for any those elements a conjugation map similar to (1.1) is an isomorphism [Sev11].

All infinite families of finite simple groups, except those of the alternating groups, consist of finite reductive groups. In 1976, Deligne and Lusztig proved that all of their characters can be obtained from the compactly supported étale cohomology of certain algebraic varieties XwGX_{w}^{G} constructed out of such groups [DL76], using a twist FF (a Frobenius morphism) and an element ww of the Weyl group. They showed that these virtual representations do not depend on the twisted conjugacy class of ww [DL76, Theorem 1.6], reducing much of the study of these Deligne-Lusztig varieties to minimal length elements. Any minimal length element is elliptic in a standard parabolic corresponding to a Levi subgroup LL, and one can show that the orbit space satisfies GF\XwGLF\XwLG^{F}\backslash X_{w}^{G}\simeq L^{F}\backslash X_{w}^{L}, thus reducing to the case where ww is elliptic and has minimal length. Lusztig and his former student He generalised the cross section (1.4) to these elements, and deduced from this that GF\XwGG^{F}\backslash X_{w}^{G} is universally homeomorphic to the quotient of 𝔸(w)\mathbb{A}^{\ell(w)} by a finite torus [HL12] (which implies a cohomology vanishing theorem [OR08], simplifying Lusztig’s classification of the representations of finite reductive groups [Lus84]). This statement was later generalised by He to obtain a dimension formula for affine Deligne-Lusztig varieties [He14], which are closely related to the reduction of integral models of Shimura varieties in arithmetic geometry. He and Lusztig mention that it

is not clear to us what is the relation of the Weyl group elements considered [by Sevostyanov] with those considered in this paper \attrib[HL12, §0.3]

and the original aim of this paper and its prequel [Mal21] was to provide a common generalisation explaining this.

Following work by Springer [Spr69], Grothendieck around 1969 obtained a simultaneous resulution of the singularities of the fibres of GG//GT/WG\rightarrow G/\!\!/G\simeq T/W and conjectured that a transverse slice to a subregular element of GG would yield a universal deformation of the corresponding Du Val-Klein singularity, and similarly for its Lie algebra. This was proven by Brieskorn [Bri71], and Slodowy thoroughly studied these deformations through suitable slices in the Lie algebra [Slo80]. Already appearing in the work of Harish-Chandra on invariant distributions on Lie algebras [HC64], Slodowy slices play a crucial rôle in the classification of Whittaker representations as the semiclassical limits of WW-algebras [Kos78, Pre02], have recently been applied to reconstruct Khovanov homology [SS06, AS19] and numerous physicists are currently using them in their study of supersymmetric gauge theories (e.g. [GW09a, GW09b]).

Meanwhile, Drinfeld showed that every reductive Lie algebra (interpreted as a Casimir Lie algebra with the Killing form) admits a unique quantisation, up to gauge transformations [Dri89]; semiclassically, this corresponds to a choice of factorisable rr-matrix, and those were classified by Belavin-Drinfeld [BD82]. In his study of zero curvature conditions for soliton equations, Semenov-Tian-Shansky discovered a natural Poisson bracket on GG associated with these rr-matrices, which is compatible with conjugation [STS85].

By using the Kazhdan-Lusztig map [KL88], Sevostyanov associated certain Weyl group elements to the subregular unipotent classes and determined that his corresponding slices have the correct dimension, yielding analogous results. He subsequently showed that after modifying the standard rr-matrix, the Semenov-Tian-Shansky bracket on GG reduces to his slices and lifts to Poisson structures on these minimal surface singularities [Sev11]. Very recently, it was shown that the analogue of the Grothendieck-Springer resolution for GG-bundles on elliptic curves [BZN15, Dav19] yields del Pezzo surfaces [GSB21, Dav20], whilst slices in the affine Grassmannian were quantised [KWWY14] and immediately applied in the study of Coulomb branches in physics (e.g. [BDG17, BFN19]).

Employing Lusztig’s recent stratification of GG by unions of sheets of conjugacy classes [Lus15], Sevostyanov verified that any conjugacy class of GG is strictly transversally intersected by one of his slices [Sev19]. This forms the main ingredient in Sevostyanov’s approach [Sev21] to the long-standing De Concini-Kac-Procesi conjecture on the dimensions of irreducible representations of quantum groups at roots of unity [DCKP92, §6.8] (the analogous Kac-Weisfeiler conjecture for Lie algebras [VK71] was proven with Slodowy slices [Pre95]). Furthermore, Sevostyanov’s slices have been used in an attempt to obtain a group analogue [CE15] of Katsylo’s theorem [Kat82], which itself has been applied e.g. to analyse the space of one-dimensional representations of finite WW-algebras and to the theory of primitive ideals [PT14].

It is natural in the theory of reductive Poisson-Lie groups and Drinfeld-Jimbo quantum groups to fix a torus and Borel subgroup as before. Recall from the introduction of [Mal21] the notion and notation for twisted finite Coxeter groups, (firmly) convex elements, reduced braids, braid power bounds, roots, stable roots, convex sets of roots, inversion sets, weak left Bruhat-Chevalley orders, right Deligne-Garside factors and normal forms and the braid equation

DG(bwd)=pb(w).\mathrm{DG}(b_{w}^{d})=\mathrm{pb}(w). (1.2)

The relevance of this equation to the main theorem is through its close relationship with

Definition 1.2.

Let ww be an element of a twisted Weyl group. For any positive root β\beta of its root system, we construct a subset of positive roots

crossw(β):={w(β+i=1mβi):β1,,βmw,m0}+.\mathrm{cross}_{w}(\beta):=\bigl{\{}w(\beta+\sum_{i=1}^{m}\beta_{i})\in\mathfrak{R}:\beta_{1},\ldots,\beta_{m}\in\mathfrak{R}_{w},m\geq 0\bigr{\}}\cap\mathfrak{R}_{+}.

This extends to a map

crossw:{subsets of +}{subsets of +},𝔑β𝔑crossw(β),\mathrm{cross}_{w}:\{\textrm{subsets of }\mathfrak{R}_{+}\}\longrightarrow\{\textrm{subsets of }\mathfrak{R}_{+}\},\qquad\mathfrak{N}\longmapsto\bigcup_{\beta\in\mathfrak{N}}\mathrm{cross}_{w}(\beta),

and we let crosswd()\mathrm{cross}_{w}^{d}(\cdot) denote its dd-th iterate for integers d0d\geq 0.

A more group-theoretic interpretation is given in Corollary 3.5, where it is explained that crossw()\mathrm{cross}_{w}(\cdot) is a (simplification of a) first-order approximation to the polynomial equations appearing in our generalisation of the cross section (1.1). In Lemma 2.6 we prove that crossw()\mathrm{cross}_{w}(\cdot) naturally extends to twisted braid monoids, which is one of the ingredients in the proof of the following

Lemma.

Let w be an element of a twisted Weyl group WW.

  1. (i)

    For any simple root α\alpha not in the inversion set w\mathfrak{R}_{w}, the set crossw(α)\mathrm{cross}_{w}(\alpha) contains simple roots.

  2. (ii)

    For any other element ww^{\prime} of WW and integer d0d\geq 0, we have

    DG(bwd)wif and only if crosswd(w)=,\mathrm{DG}(b_{w}^{d})\geq w^{\prime}\qquad\textrm{if and only if }\qquad\mathrm{cross}_{w}^{d}(\mathfrak{R}_{w^{\prime}})=\varnothing,

    if and only if crosswd(w)\mathrm{cross}_{w}^{d}(\mathfrak{R}_{w^{\prime}}) does not contain any simple roots.

  3. (iii)

    Moreover, for any d>|+\stw|(w)d>|\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}|-\ell(w) we have

    w is convex and satisfies (1.2)if and only if crosswd(+\stw)=,w\textrm{ is convex and satisfies }(\ref{eq:braid-equation})\qquad\textrm{if and only if }\qquad\mathrm{cross}_{w}^{d}(\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w})=\varnothing,

    if and only if crosswd(+\stw)\mathrm{cross}_{w}^{d}(\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}) does not contain any simple roots.

Example 1.3.

Consider w=s2w=s_{2} in type 𝖡2\mathsf{B}_{2}, then the only simple root in crossw(α1)\mathrm{cross}_{w}(\alpha_{1}) is α1=w(α1+2α2)\alpha_{1}=w(\alpha_{1}+2\alpha_{2}).

Example 1.4.

Consider w=s1s3w=s_{1}s_{3} in type 𝖠3\mathsf{A}_{3}, then the only simple root in crossw(α2)\mathrm{cross}_{w}(\alpha_{2}) is α2=w(α2+α1+α3)\alpha_{2}=w(\alpha_{2}+\alpha_{1}+\alpha_{3}).

Our proof of transversality involves root combinatorics which through this lemma is closely intertwined with

Corollary A.

Let ww be a convex element of a twisted Weyl group. Then

DG(bwd)=pb(w)if and only if DG(bw1d)=pb(w).\mathrm{DG}(b_{w}^{d})=\mathrm{pb}(w)\qquad\textrm{if and only if }\qquad\mathrm{DG}(b_{w^{-1}}^{d})=\mathrm{pb}(w).

Typically however, these Deligne-Garside factors are not very similar:

Example 1.5.

Consider w=s1s2s3s1s2w=s_{1}s_{2}s_{3}s_{1}s_{2} in type 𝖡3\mathsf{B}_{3}; it is convex and for any integer d>1d>1 we have

DGN(bwd)=bwdandDGN(bw1d)=bw1s1bw1d2bs1w1.\mathrm{DGN}(b_{w}^{d})=b_{w}^{d}\qquad\textrm{and}\qquad\mathrm{DGN}(b_{w^{-1}}^{d})=b_{w^{-1}s_{1}}^{\,}b_{w^{-1}}^{d-2}b_{s_{1}w^{-1}}^{\,}.

The perspective furnished by Definition 1.2 allows us to construct cross sections out of quite general data:

Definition 1.6.

Let ww be an element of a twisted Weyl group and let 𝔑+\stw\mathfrak{N}\subseteq\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w} be a convex subset of positive roots. Then we will say that 𝔑\mathfrak{N} is (ww-)nimble if it contains w\mathfrak{R}_{w} and furthermore

w(𝔑\w)𝔑.w(\mathfrak{N}\backslash\mathfrak{R}_{w})\subseteq\mathfrak{N}.
Example 1.7.

From [Mal21, Proposition 4.21] it follows that the set +\stw\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w} is nimble itself if and only if the element ww is convex.

Notation 1.8.

Given two subsets of roots 𝔑,𝔏\mathfrak{N},\mathfrak{L} of a root system \mathfrak{R}, we write

𝔑+𝔏:={c0β0+c1β1:β0𝔑,β1𝔏,c0,c1>0}.\mathfrak{N}+\mathfrak{L}:=\{c_{0}\beta_{0}+c_{1}\beta_{1}\in\mathfrak{R}:\beta_{0}\in\mathfrak{N},\beta_{1}\in\mathfrak{L},c_{0},c_{1}\in\mathbb{R}_{>0}\}.
Definition 1.9.

Let ww be an element of a twisted Weyl group and let 𝔑\mathfrak{N} a ww-nimble set. We will say that a convex subset of roots 𝔏\𝔑\mathfrak{L}\subseteq\mathfrak{R}\backslash\mathfrak{N} is a (𝔑\mathfrak{N}-)leavener when

  1. (i)

    w(𝔏)=𝔏=𝔏w(\mathfrak{L})=\mathfrak{L}=-\mathfrak{L},

  2. (ii)

    the set w𝔏\mathfrak{R}_{w}\sqcup\mathfrak{L} is convex, and

  3. (iii)

    the set 𝔑𝔏\mathfrak{N}\sqcup\mathfrak{L} is convex.

We will then refer to (𝔑,𝔏)(\mathfrak{N},\mathfrak{L}) as a (ww-)crossing pair. Note that (i) implies w1𝔏=\mathfrak{R}_{w^{-1}}\cap\mathfrak{L}=\varnothing. If furthermore +𝔑𝔏\mathfrak{R}_{+}\subseteq\mathfrak{N}\sqcup\mathfrak{L} (forcing 𝔏\mathfrak{L} to be a standard parabolic subsystem), then we will say that it is a (ww-)slicing pair.

Proposition 2.21 explains that crossing pairs arise quite naturally. In particular:

Proposition.

Let ww be an element of a twisted Weyl group. If ww is convex, then for any convex subset 𝔏stw\mathfrak{L}\subseteq\mathfrak{R}_{\mathrm{st}}^{w} satisfying w(𝔏)=𝔏=𝔏w(\mathfrak{L})=\mathfrak{L}=-\mathfrak{L}, both

(+\stw,𝔏)and(DG(bwd),𝔏)(\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w},\mathfrak{L})\qquad\textrm{and}\qquad(\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})},\mathfrak{L})

for any d1d\geq 1 are crossing pairs.

I expect that the most important class of cross sections will be the one arising from firmly convex elements with crossing pair (+\w,w)(\mathfrak{R}_{+}\backslash\mathfrak{R}^{w},\mathfrak{R}^{w}). In order to extend our results to twisted conjugacy classes and make the proofs slightly cleaner, we will employ the language of twisted reductive groups:

Definition 1.10.

Let G~\tilde{G} be a split reductive group over a scheme with chosen Borel subgroup, and let Ω\Omega be a group of twists of its Weyl group W~\tilde{W}; these extend to automorphisms of G~\tilde{G}. Also let ϕ\phi be an automorphism of the base scheme. We will then call G:=ϕΩG~G:=\phi\Omega\ltimes\tilde{G} a twisted split reductive group, and W:=ΩW~W:=\Omega\ltimes\tilde{W} its Weyl group.

Notation 1.11.

Let ww be an element of its Weyl group and let (𝔑,𝔏)(\mathfrak{N},\mathfrak{L}) be a ww-crossing pair. Then we let NN+N\subseteq N_{+} denote the unipotent subgroup corresponding to the roots in 𝔑\mathfrak{N}, and let LL denote the reductive subgroup of GG generated by a chosen subgroup TT^{\prime} of the torus TT, and by the root subgroups corresponding to 𝔏\mathfrak{L}.

In the firmly convex case, one typically sets TT^{\prime} to be the set of fixed points TwT^{w} under the action of ww on the torus TT; we denote its Lie algebra by 𝔱w\mathfrak{t}^{w}.

Notation 1.12.

Factorisable rr-matrices are parametrised by a Belavin-Drinfeld triple 𝔗=(𝔗0,𝔗1,τ)\mathfrak{T}=(\mathfrak{T}_{0},\mathfrak{T}_{1},\tau) defining a nilpotent isomorphism τ:𝔗0𝔗1\tau:\mathfrak{T}_{0}\rightarrow\mathfrak{T}_{1} between two subsets of simple roots, and an element r0r_{0} in 𝔱𝔗𝔱𝔗\mathfrak{t}_{\mathfrak{T}}\wedge\mathfrak{t}_{\mathfrak{T}}, where 𝔱𝔗\mathfrak{t}_{\mathfrak{T}} is a certain subspace of the Lie algebra of the maximal torus TT. Given such a triple 𝔗\mathfrak{T} and a set of roots 𝔏\mathfrak{L}, by 𝔗𝔏\mathfrak{T}\subseteq\mathfrak{L} we mean that 𝔗0\mathfrak{T}_{0} and 𝔗1\mathfrak{T}_{1} are both contained in 𝔏\mathfrak{L}; thus this is always satisfied for the empty triple.

The main aim of this paper is to generalise He-Lusztig’s, Sevostyanov’s and Steinberg’s slices to the following

Theorem.

Let GG be a twisted split reductive group over a scheme with chosen Borel subgroup containing a maximal torus TT. Pick any element ww in its Weyl group and let (𝔑,𝔏)(\mathfrak{N},\mathfrak{L}) be a ww-crossing pair such that for some integer d0d\geq 0, or equivalently for any integer d>|𝔑|(w)d>|\mathfrak{N}|-\ell(w), there is an identity

crosswd(𝔑)=.\mathrm{cross}_{w}^{d}(\mathfrak{N})=\varnothing. (1.3)
  1. (i)

    Then the (right) conjugation map

    N×w˙LNwNw˙LN,(n,g)n1gnN\times\dot{w}LN_{w}\longrightarrow N\dot{w}LN,\qquad(n,g)\longmapsto n^{-1}gn (1.4)

    is an isomorphism.

  2. (ii)

    If furthermore LL contains TwT^{w} and the pair is slicing, then the conjugation map

    G×w˙LNwG,(n,g)n1gnG\times\dot{w}LN_{w}\longrightarrow G,\qquad(n,g)\longmapsto n^{-1}gn (1.5)

    is smooth; in other words,

    the subspace w˙LNw of G transversely intersects the conjugation orbits of G~.\textrm{the subspace }\dot{w}LN_{w}\textrm{ of }G\textrm{ transversely intersects the conjugation orbits of }\tilde{G}.
  3. (iii)

    If ww is firmly closed and (𝔑,𝔏)=(+\w,w)(\mathfrak{N},\mathfrak{L})=(\mathfrak{R}_{+}\backslash\mathfrak{R}^{w},\mathfrak{R}^{w}) and LTTwL\cap T\subseteq T^{w}, then the Semenov-Tian-Shansky bracket associated to a factorisable rr-matrix with Belavin-Drinfeld triple 𝔗\mathfrak{T} satisfying 𝔗𝔏\mathfrak{T}\subseteq\mathfrak{L} reduces to the subspace w˙LNw\dot{w}LN_{w} if and only if

    1. (a)

      LT=TwL\cap T=T^{w},

    2. (b)

      r0r_{0} preserves 𝔱w\mathfrak{t}^{w}, and

    3. (c)

      r0r_{0} acts on its orthogonal complement as the Cayley transform

      1+w1w.\frac{1+w}{1-w}.

My reasons for choosing this formulation are given at the start of section §3.

Example 1.13.

Let G=SL3G=\mathrm{SL}_{3}, and pick the Borel subgroup and maximal torus and of upper triangular matrices and diagonal matrices. Set w:=s1s2s1w:=s_{1}s_{2}s_{1} and lift it to

w˙=[001010100]G.\dot{w}=\begin{bmatrix}0&0&1\\ 0&-1&0\\ 1&0&0\end{bmatrix}\in G.

As w=ww=w_{\circ}, part (i) of the main Theorem implies that the conjugation map sending

([1n1n1201n2001],[00t0t2x2tx1x12])N+×w˙TwN+\left(\begin{bmatrix}1&n_{1}&n_{12}\\ 0&1&n_{2}\\ 0&0&1\end{bmatrix},\begin{bmatrix}0&0&t\\ 0&-t^{-2}&x_{2}\\ t&x_{1}&x_{12}\end{bmatrix}\right)\in N_{+}\times\dot{w}T^{w}N_{+}

to

[(n1n2n12)tn1t2+(n1n2n12)(n1t+x1)n1n2t2+(n1n2n12)(n12t+x12+n2x1)+tn1x2n2tt2n1n2tn2x1x2t2n2(1+t2(n12t+n2x1+x12))tx1+n1tn12t+n2x1+x12]\begin{bmatrix}(n_{1}n_{2}-n_{12})t&n_{1}t^{-2}+(n_{1}n_{2}-n_{12})(n_{1}t+x_{1})&n_{1}n_{2}t^{-2}+(n_{1}n_{2}-n_{12})(n_{12}t+x_{12}+n_{2}x_{1})+t-n_{1}x_{2}\\ -n_{2}t&-t^{-2}-n_{1}n_{2}t-n_{2}x_{1}&x_{2}-t^{-2}n_{2}(1+t^{2}(n_{12}t+n_{2}x_{1}+x_{12}))\\ t&x_{1}+n_{1}t&n_{12}t+n_{2}x_{1}+x_{12}\end{bmatrix}

is an isomorphism onto Nw˙TwNN\dot{w}T^{w}N. Over a field the subregular conjugacy classes of GG can be parametrised by

{[00c20c220c20c1]:c1,c2𝒪S}w˙TwN+,\left\{\begin{bmatrix}0&0&c_{2}\\ 0&-c_{2}^{-2}&0\\ c_{2}&0&c_{1}\end{bmatrix}:c_{1},c_{2}\in\mathcal{O}_{S}\right\}\subset\dot{w}T^{w}N_{+},

and from part (ii) of the main Theorem and a dimension count it now follows that they are strictly transversally intersected by the subspace w˙TwN+\dot{w}T^{w}N_{+}.

Example 1.14.

Consider w=s1s2s3s1s2w=s_{1}s_{2}s_{3}s_{1}s_{2} in type 𝖡3\mathsf{B}_{3} again and let 𝔑:=w{α1233}\mathfrak{N}:=\mathfrak{R}_{w}\sqcup\{\alpha_{1233}\}. This set is convex and although DG(bwi)=bwi\mathrm{DG}(b_{w}^{i})=b_{w}^{i}, we have w(α1233)ww(\alpha_{1233})\in\mathfrak{R}_{w} so 𝔑\mathfrak{N} is nimble and hence the cross section holds.

Example 1.15.

Consider Spaltenstein’s example of the (inverse of the) element w=s1s2s3s4s5s3s4s1s2w=s_{1}s_{2}s_{3}s_{4}s_{5}s_{3}s_{4}s_{1}s_{2} in type 𝖠5\mathsf{A}_{5} with the conjugation map corresponding to the crossing pair (𝔑,𝔏)=(+,)(\mathfrak{N},\mathfrak{L})=(\mathfrak{R}_{+},\varnothing); by induction on i0i\geq 0 one can show that

DGN(bwi+2)=bws5b451234312ibs5w,\mathrm{DGN}(b_{w}^{i+2})=b_{ws_{5}}^{\,}b_{451234312}^{i}b_{s_{5}w}^{\,},

so as ww is elliptic the inequality s5ww=pb(w)s_{5}w\neq w_{\circ}=\mathrm{pb}(w) and main Lemma imply that we cannot invoke the main Theorem here. In fact, (some of) such Coxeter elements of length 9 in 𝖠5\mathsf{A}_{5} are the “smallest” elliptic examples in type 𝖠\mathsf{A} such that the braid equation (1.2) does not hold. One computes that

crosswi+2(+)=crossw2(+)={α23,α2345,α3,α345,α5}\mathrm{cross}_{w}^{i+2}(\mathfrak{R}_{+})=\mathrm{cross}_{w}^{2}(\mathfrak{R}_{+})=\{\alpha_{23},\alpha_{2345},\alpha_{3},\alpha_{345},\alpha_{5}\} (1.6)

and as crossw()\mathrm{cross}_{w}(\cdot) can be interpreted as a first-order approximation to the equations of (1.4) (after parametrising both sides into products of root subgroups, see Corollary 3.5), this suggests that there might be a linear relationship. Indeed, employing the usual lift

w˙=[000001001000100000000010010000000100]\dot{w}=\begin{bmatrix}0&0&0&0&0&1\\ 0&0&-1&0&0&0\\ 1&0&0&0&0&0\\ 0&0&0&0&-1&0\\ 0&-1&0&0&0&0\\ 0&0&0&1&0&0\end{bmatrix}

and studying the resulting equations for the root subgroups of the roots in (1.6), one obtains a family of first-order counterexamples which exponentiates to (a slight generalisation of) Spaltenstein’s counterexample

{([100000010st0s001s0st1000100000010000001],[00000100100010t100000001001t0000001tt1])N+×w˙Nw:s,t𝒪S}.\left\{\left(\begin{bmatrix}1&0&0&0&0&0\\ 0&1&0&-st&0&-s\\ 0&0&1&s&0&st^{-1}\\ 0&0&0&1&0&0\\ 0&0&0&0&1&0\\ 0&0&0&0&0&1\end{bmatrix},\begin{bmatrix}0&0&0&0&0&1\\ 0&0&-1&0&0&0\\ 1&0&t^{-1}&0&0&0\\ 0&0&0&0&-1&0\\ 0&-1&-t&0&0&0\\ 0&0&0&1&t&t^{-1}\end{bmatrix}\right)\in N_{+}\times\dot{w}N_{w}:s,t\in\mathcal{O}_{S}\right\}.

Nevertheless, the converse to (i) does not hold, as linear combinations of the polynomials in the first-order approximation can sometimes cancel each other out:

Example 1.16.

Consider the Coxeter element w=s2s3s4s5s6s1s2s3s4s5s3s2w=s_{2}s_{3}s_{4}s_{5}s_{6}s_{1}s_{2}s_{3}s_{4}s_{5}s_{3}s_{2} in type 𝖠6\mathsf{A}_{6}. Then bwib_{w}^{i} is in Deligne-Garside normal form for any i0i\geq 0 so from part (iii) of the main Lemma it again follows that equation (1.3) is never satisfied, but a (rather lengthy) calculation shows that the conjugation map (1.4) corresponding to the slicing pair (𝔑,𝔏)=(+,)(\mathfrak{N},\mathfrak{L})=(\mathfrak{R}_{+},\varnothing) is an isomorphism.

Finally, Sevostyanov and He-Lusztig proved the cross section isomorphism (1.4) for certain firmly convex elements ww with slicing pair (𝔑,𝔏)=(+\w,w)(\mathfrak{N},\mathfrak{L})=(\mathfrak{R}_{+}\backslash\mathfrak{R}^{w},\mathfrak{R}^{w}), under some extra conditions on the base ring. The relationship between the elements they consider is explained in the prequel [Mal21, §2.1], and it is proven there that all of these indeed satisfy the braid equation (1.2) when d>|+\w|(w)d>|\mathfrak{R}_{+}\backslash\mathfrak{R}^{w}|-\ell(w) [Mal21, Theorem B]; there are many convex (and firmly convex) elements satisfying (1.2) which are contained are in neither, yielding more transverse slices. The main conclusion of these two papers is now obtained by combining their main theorems with Sevostyanov’s dimension calculations [Sev19] to

Corollary B.

The Weyl group elements considered by He-Lusztig and the elements Sevostyanov uses to construct strictly transverse slices (for connected reductive groups over algebraically closed fields) are all minimally dominant, and conversely all minimally dominant elements of conjugacy classes appearing in Lusztig’s partition furnish strictly transverse slices with natural Poisson brackets.

Acknowledgements

I thank Ian Grojnowski for comments and I am grateful to Dominic Joyce, Balázs Szendrői and the Mathematical Institute of the University of Oxford for an excellent stay, where some of this paper was written. This visit was supported by the Centre for Quantum Geometry of Moduli Spaces at Aarhus University and the Danish National Research Foundation. This work was also supported by EPSRC grant EP/R045038/1.

2 Crossing roots

A common strategy, dating back to Killing’s work around 1889, is to study reductive groups and Weyl groups through their root systems; we follow this perspective by developing properties of crossw()\mathrm{cross}_{w}(\cdot). Throughout this paper, root systems will always be assumed to be crystallographic; we begin by proving a lemma on decomposing sums of roots, that we will employ several times:

Definition 2.1.

We will say that a sequence (β1,,βm)(\beta_{1},\ldots,\beta_{m}) of roots in a crystallographic root system is a summing sequence if each of the partial sums i=1jβi\sum_{i=1}^{j}\beta_{i} is a root for 1jm1\leq j\leq m, and if their total sum is denoted by γ:=i=1mβi\gamma:=\sum_{i=1}^{m}\beta_{i} then we will denote this sequence as

(β1,,βm)=γ.\sum(\beta_{1},\ldots,\beta_{m})=\gamma.

If furthermore each of these roots βi\beta_{i} lies in a subset of roots 𝔑\mathfrak{N}, then we will call this a summing sequence in 𝔑\mathfrak{N}.

The following lemma shows how summing sequences may be constructed:

Lemma 2.2.

Let β0,,βm\beta_{0},\ldots,\beta_{m} be roots in a crystallographic root system such that their sum i=0mβi\sum_{i=0}^{m}\beta_{i} is also a root.

  1. (i)

    [Bou68, Proposition VI.1.19] Amongst these roots there exists a root βj\beta_{j} such that the difference

    i=0mβiβj\sum_{i=0}^{m}\beta_{i}-\beta_{j}

    is either a root or is zero; it is strictly positive when m>0m>0 and each root βi\beta_{i} is positive.

  2. (ii)

    Suppose briefly that m=3m=3 and that βi+βj0\beta_{i}+\beta_{j}\neq 0 for each distinct pair i,j{1,2,3}i,j\in\{1,2,3\}. Then at least two of the three sums

    β1+β2,β1+β3,β2+β3\beta_{1}+\beta_{2},\qquad\beta_{1}+\beta_{3},\qquad\beta_{2}+\beta_{3}

    are also roots.

  3. (iii)

    Hence we may obtain from these roots β1,,βm\beta_{1},\ldots,\beta_{m} by reordering and deleting a summing sequence

    (βi1,,βim)=i=1mβi.\sum(\beta_{i_{1}},\ldots,\beta_{i_{m^{\prime}}})=\sum_{i=1}^{m}\beta_{i}.

    If each of the roots βi\beta_{i} is positive, then the resulting sequence βi1,,βim\beta_{i_{1}},\ldots,\beta_{i_{m}} is simply a reordering and we may choose the sequence to start with any of the βi\beta_{i}.

In particular, if each of the βi\beta_{i} lie in a convex subset 𝔑\mathfrak{N}, then i=0mβi\sum_{i=0}^{m}\beta_{i} also lies in 𝔑\mathfrak{N}.

Proof.

(i): We follow the proof of [Bou68, Proposition VI.1.19]; let’s write γ:=i=1mβi\gamma:=\sum_{i=1}^{m}\beta_{i}. If (βi,γ)0(\beta_{i},\gamma)\leq 0 for each root βi\beta_{i}, then by linearity also (γ,γ)=(i=1mβi,γ)0(\gamma,\gamma)=(\sum_{i=1}^{m}\beta_{i},\gamma)\leq 0 which contradicts that γ\gamma is a root. Thus there must be an inequality (βj,γ)>0(\beta_{j},\gamma)>0 for some jj, which in a crystallographic root system implies that γβj\gamma-\beta_{j} is either a root or is zero.

(ii): By (i), we may assume after relabelling that β1+β2\beta_{1}+\beta_{2} is a root (as it is not zero by assumption). If (βi,β1+β2)0(\beta_{i},\beta_{1}+\beta_{2})\leq 0 for both i{1,2}i\in\{1,2\} then we derive the same contradiction

(β1+β2,β1+β2)0(\beta_{1}+\beta_{2},\beta_{1}+\beta_{2})\leq 0

as before, so we may assume that say (β1,β1+β2)>0(\beta_{1},\beta_{1}+\beta_{2})>0. If β1+β3\beta_{1}+\beta_{3} is a root the claim follows and if it is not then (β1,β3)0(\beta_{1},\beta_{3})\geq 0. This would yield

(β1,β1+β2+β3)=(β1,β1+β2)+(β1,β3)>0,(\beta_{1},\beta_{1}+\beta_{2}+\beta_{3})=(\beta_{1},\beta_{1}+\beta_{2})+(\beta_{1},\beta_{3})>0,

which implies that (β1+β2+β3)β1=β2+β3(\beta_{1}+\beta_{2}+\beta_{3})-\beta_{1}=\beta_{2}+\beta_{3} is a root.

(iii): The first claim of (iii) follows from (i) by induction on mm. As a nontrivial sum of positive roots is never zero the sequence must be a reordering if each of the roots βi\beta_{i} are positive. For the final claim we induct on mm, so we may assume that βim\beta_{i_{m}} is the chosen root we want the sequence to start with. Writing γ<m1:=j=1m2βij\gamma_{<m-1}:=\sum_{j=1}^{m-2}\beta_{i_{j}}, part (ii) yields that at least one of

γm1+βimorβim1+βim\gamma_{m-1}+\beta_{i_{m}}\qquad\textrm{or}\qquad\beta_{i_{m-1}}+\beta_{i_{m}}

is a root. In the former case we apply the induction hypothesis to that sum to conclude, and in the latter case we apply the induction hypothesis to the set of roots βi1,,βim2,βim1+βim\beta_{i_{1}},\ldots,\beta_{i_{m-2}},\beta_{i_{m-1}}+\beta_{i_{m}} to find a summing sequence beginning with βim1+βim\beta_{i_{m-1}}+\beta_{i_{m}}, which immediately yields one starting with βim\beta_{i_{m}}. ∎

Example 2.3.

Let β,γ\beta,\gamma be a pair of roots in a root system such that neither β+γ\beta+\gamma nor βγ\beta-\gamma are roots (e.g. γ=β\gamma=\beta), then

β+γ+(γ)\beta+\gamma+(-\gamma)

is a root but none of the three sums in (ii) are roots.

2.1 Crossing for braids

In this subsection we continue developing results on decomposing sums of roots, and as a first application we show that the definition of crossw()\mathrm{cross}_{w}(\cdot) extends to the braid monoid. We will give another proof of this using root subgroups in Proposition 3.6; that proof is arguably more intuitive, but since the main Lemma and its various corollaries might be of interest independent of reductive groups I decided to include a proof using only roots.

Part (ii) and (iii) of the previous lemma yield the following corollary:

Corollary 2.4.

Suppose 𝔑,𝔑\mathfrak{N},\mathfrak{N}^{\prime} are two subsets of roots of a crystallographic root system, such that 𝔑,𝔑\mathfrak{N},\mathfrak{N}^{\prime} and also their union 𝔑𝔑\mathfrak{N}\cup\mathfrak{N}^{\prime} are convex. Suppose furthermore that β\beta is a root such that β-\beta does not lie in 𝔑𝔑\mathfrak{N}\cup\mathfrak{N}^{\prime}. Then from any summing sequence

(β,β~1,,β~m)=:γ\sum(\beta,\tilde{\beta}_{1},\ldots,\tilde{\beta}_{m})=:\gamma

with β~i𝔑𝔑\tilde{\beta}_{i}\in\mathfrak{N}\cup\mathfrak{N}^{\prime} for each ii, we can obtain one

(β,β1,,βk,β1,,βk)=γ\sum(\beta,\beta_{1},\ldots,\beta_{k},\beta_{1}^{\prime},\ldots,\beta_{k^{\prime}}^{\prime})=\gamma

with βi𝔑\beta_{i}\in\mathfrak{N} and βi𝔑\beta_{i}^{\prime}\in\mathfrak{N}^{\prime}, each of them obtained from the β~i\tilde{\beta}_{i} by rearranging and summing some of them.

Proof.

We induct on mm and within that we induct on 2lm2\leq l\leq m, which is the first time β~l\tilde{\beta}_{l} lies in 𝔑\mathfrak{N} but β~l1\tilde{\beta}_{l-1} lies in 𝔑\mathfrak{N}^{\prime}; if there is no such ll we are already done. Set γ<l1:=β+i=1l2β~i\gamma_{<l-1}:=\beta+\sum_{i=1}^{l-2}\tilde{\beta}_{i} and consider the sum

γ<l1+β~l1+β~l.\gamma_{<l-1}+\tilde{\beta}_{l-1}+\tilde{\beta}_{l}.

If either γ<l1=0\gamma_{<l-1}=0 or β~l1+β~l=0\tilde{\beta}_{l-1}+\tilde{\beta}_{l}=0 then we may shorten the sequence and conclude from the induction hypothesis on mm. Moreover if

β+i=1l2β~i+β~l1=γ<l1+β~l1=0\beta+\sum_{i=1}^{l-2}\tilde{\beta}_{i}+\tilde{\beta}_{l-1}=\gamma_{<l-1}+\tilde{\beta}_{l-1}=0

then rewriting yields β=i=1l2β~i+β~l1-\beta=\sum_{i=1}^{l-2}\tilde{\beta}_{i}+\tilde{\beta}_{l-1}. From the last statement in part (iii) it would follow that β-\beta lies in the convex set 𝔑𝔑\mathfrak{N}\cup\mathfrak{N}^{\prime}. The case γ<l1+β~l=0\gamma_{<l-1}+\tilde{\beta}_{l}=0 similarly yields a contradiction. Hence by part (ii) at least one of γ<l1+β~l\gamma_{<l-1}+\tilde{\beta}_{l} or β~l1+β~l\tilde{\beta}_{l-1}+\tilde{\beta}_{l} is a root; in the latter case it lies in 𝔑𝔑\mathfrak{N}\cup\mathfrak{N}^{\prime} and we may shorten the sequence, whereas in the former case we may now swap the roots β~l1\tilde{\beta}_{l-1} and β~l\tilde{\beta}_{l} in the sequence to lower ll and apply the induction hypothesis on ll. ∎

Definition 2.5.

Let b:=bw¯:=bwdbw1b:=b_{\underline{w}}:=b_{w_{d}}\cdots b_{w_{1}} be a braid in a braid monoid, constructed out of a sequence of elements w¯=(wd,,w1)\underline{w}=(w_{d},\ldots,w_{1}) lying in the corresponding twisted Weyl group. Given a positive root β\beta or a subset of positive roots 𝔑+\mathfrak{N}\subseteq\mathfrak{R}_{+}, construct the set of roots

crossb(β):=crosswdcrossw2crossw1(β)\mathrm{cross}_{b}(\beta):=\mathrm{cross}_{w_{d}}\cdots\mathrm{cross}_{w_{2}}\mathrm{cross}_{w_{1}}(\beta)

and extend it to a map

crossb:{subsets of +}{subsets of +},𝔑β𝔑crossb(β).\mathrm{cross}_{b}:\{\textrm{subsets of }\mathfrak{R}_{+}\}\longrightarrow\{\textrm{subsets of }\mathfrak{R}_{+}\},\qquad\mathfrak{N}\longmapsto\bigcup_{\beta\in\mathfrak{N}}\mathrm{cross}_{b}(\beta).

We let crossw¯(γ,β)\mathrm{cross}_{\underline{w}}(\gamma,\beta) denote the set of sequences of elements in the wi\mathfrak{R}_{w_{i}} “confirming” that γ\gamma lies in crossb(β)\mathrm{cross}_{b}(\beta); more precisely, it is the set of subsets of roots

𝔑m××𝔑2×𝔑1wm××w2×w1\mathfrak{N}_{m}\times\cdots\times\mathfrak{N}_{2}\times\mathfrak{N}_{1}\subseteq\mathfrak{R}_{w_{m}}\times\cdots\times\mathfrak{R}_{w_{2}}\times\mathfrak{R}_{w_{1}}

with the property that the corresponding sequence β=β0,,βd\beta=\beta_{0}^{\prime},\ldots,\beta_{d}^{\prime} inductively constructed via

βj:=wj(βj1+β~𝔑jβ~)\beta_{j}^{\prime}:=w_{j}(\beta_{j-1}^{\prime}+\sum_{\tilde{\beta}\in\mathfrak{\mathfrak{N}}_{j}}\tilde{\beta})

for 1jd1\leq j\leq d, consists solely of roots which are all positive, and satisfies βd=γ\beta_{d}^{\prime}=\gamma.

Analysing these sequences shows that the set of roots crossb()\mathrm{cross}_{b}(\cdot) is well-defined:

Lemma 2.6.

Let w¯,w¯\underline{w},\underline{w}^{\prime} be two sequences of elements in a twisted Weyl group such that there is an equality bw¯=bw¯b_{\underline{w}}=b_{\underline{w}^{\prime}} in the associated braid monoid, and let β,γ\beta,\gamma be positive roots in its root system. One can non-canonically construct “transfer maps”

crossw¯(γ,β)crossw¯(γ,β),\mathrm{cross}_{\underline{w}}(\gamma,\beta)\rightleftarrows\mathrm{cross}_{\underline{w}^{\prime}}(\gamma,\beta),

mapping nontrivial sequences to nontrivial sequences.

In particular, for any subset of positive roots 𝔑\mathfrak{N} and any braid bb in the corresponding braid monoid, the set of roots crossb(𝔑)\mathrm{cross}_{b}(\mathfrak{N}) does not depend on the chosen decomposition of bb into reduced braids.

Proof.

We first show the claim for a reduced decomposition w=xyw=xy of an arbitrary element ww in the twisted Weyl group. A (nontrivial) sequence in cross(x,y)(γ,β)\mathrm{cross}_{(x,y)}(\gamma,\beta) rewrites as

γ=x(y(β+βyβ)+β′′xβ′′)=w(β+βyβ+y1(β′′xβ′′)),\gamma=x\bigl{(}y(\beta+\sum_{\beta^{\prime}\in\mathfrak{R}_{y}}\beta^{\prime})+\sum_{\beta^{\prime\prime}\in\mathfrak{R}_{x}}\beta^{\prime\prime}\bigr{)}=w\bigl{(}\beta+\sum_{\beta^{\prime}\in\mathfrak{R}_{y}}\beta^{\prime}+y^{-1}(\sum_{\beta^{\prime\prime}\in\mathfrak{R}_{x}}\beta^{\prime\prime})\bigr{)}, (2.1)

which through the identity

w=y1(x)y\mathfrak{R}_{w}=y^{-1}(\mathfrak{R}_{x})\sqcup\mathfrak{R}_{y} (2.2)

immediately yields a (nontrivial) sequence in crossw(γ,β)\mathrm{cross}_{w}(\gamma,\beta). Conversely, given roots β1,,βm\beta_{1},\ldots,\beta_{m} in w\mathfrak{R}_{w} such that

β+i=1mβi=w1(γ),\beta+\sum_{i=1}^{m}\beta_{i}=w^{-1}(\gamma),

obtain through part (iii) of the previous lemma a summing sequence starting with β\beta, and according to equation (2.2) each of the subsequent roots lies in one of the convex sets v1(u)v^{-1}(\mathfrak{R}_{u}) or v\mathfrak{R}_{v}. By the previous corollary we may modify the summing sequence to one which starts with β\beta, is then followed by roots in v\mathfrak{R}_{v}, and then in v1(u)v^{-1}(\mathfrak{R}_{u}). In the process, roots are only reordered or summed, so since both v1(u)v^{-1}(\mathfrak{R}_{u}) and v\mathfrak{R}_{v} consist of positive roots and the sum of positive roots is positive, it follows through (2.1) that a nontrivial sequence in crossw(γ,β)\mathrm{cross}_{w}(\gamma,\beta) yields another nontrivial sequence in cross(u,v)(γ,β)\mathrm{cross}_{(u,v)}(\gamma,\beta).

Now let w¯\underline{w} and w¯\underline{w}^{\prime} be as in the statement of this lemma. We may decompose both bw¯b_{\underline{w}} and bw¯b_{\underline{w}^{\prime}} into a product of elementary braids bib_{i} of length one and twists; one first verifies that

crossδx=crossδcrossx=crossδxδ1crossδ\mathrm{cross}_{\delta x}=\mathrm{cross}_{\delta}\mathrm{cross}_{x}=\mathrm{cross}_{\delta x\delta^{-1}}\mathrm{cross}_{\delta^{\,}}

for any twist δ\delta and element xx in the underlying untwisted Weyl group, so we may move all of the twists to the left and combine them. The equality bw¯=bw¯b_{\underline{w}}=b_{\underline{w}^{\prime}} then implies that these twists agree and can be safely ignored. This braid identity on the untwisted part implies that the first sequence transforms into the second one, through a finite sequence of braid moves sisjsi=sjsisjs_{i}s_{j}s_{i}\cdots=s_{j}s_{i}s_{j}\cdots. Now applying the result in the first paragraph several times for each such braid move, we obtain transfer maps

cross(si,sj,si,)(γ,β)crosssisjsi(γ,β)=crosssjsisj(γ,β)cross(sj,si,sj,)(γ,β)\mathrm{cross}_{(s_{i},s_{j},s_{i},\ldots)}(\gamma,\beta)\rightleftarrows\mathrm{cross}_{s_{i}s_{j}s_{i}\cdots}(\gamma,\beta)=\mathrm{cross}_{s_{j}s_{i}s_{j}\cdots}(\gamma,\beta)\rightleftarrows\mathrm{cross}_{(s_{j},s_{i},s_{j},\ldots)}(\gamma,\beta)

for any positive root β\beta. Hence we by induction on the number of braid moves to be made, we obtain transfer maps crossw¯(γ,β)crossw¯(γ,β)\mathrm{cross}_{\underline{w}}(\gamma,\beta)\rightleftarrows\mathrm{cross}_{\underline{w}^{\prime}}(\gamma,\beta). ∎

Transferring nontrivial sequences will play a key rôle in the proof of Lemma 2.19.

2.2 Crossing simple roots

In this subsection we prove part (i) of the main Lemma. It is independent of the previous subsection and in contrast, I do not know of a simple interpretation or proof in terms of root subgroups.

Proposition 2.7.

Let β0,β1,γ0,γ1\beta_{0},\beta_{1},\gamma_{0},\gamma_{1} be positive roots (or zero) in a crystallographic root system, such that their sums yield an equality

β0+β1=γ0+γ1\beta_{0}+\beta_{1}=\gamma_{0}+\gamma_{1}

between two positive roots. Then there exists a pair of indices i,j{0,1}i,j\in\{0,1\} such that

βiγj\beta_{i}-\gamma_{j}

is either a positive root or is zero.

Proof.

Since (β0+β1,γ0+γ1)>0(\beta_{0}+\beta_{1},\gamma_{0}+\gamma_{1})>0, we must have (β0+β1,γj)>0(\beta_{0}+\beta_{1},\gamma_{j})>0 for some j{0,1}j\in\{0,1\}, and thus (βi,γj)>0(\beta_{i},\gamma_{j})>0 for some i{0,1}i\in\{0,1\}. In a crystallographic root system it follows that βiγj\beta_{i}-\gamma_{j} must be either a root or zero, so if it is not a negative root we are done. If it is negative, then the complementary pair of indices i:=1ii^{\prime}:=1-i and j:=1jj^{\prime}:=1-j yields a positive root

βiγj=(βiγj).\beta_{i^{\prime}}-\gamma_{j^{\prime}}=-(\beta_{i}-\gamma_{j}).\qed
Corollary 2.8.

Let β0,,βm\beta_{0},\ldots,\beta_{m} for m0m\geq 0 and γ0,γ1\gamma_{0},\gamma_{1} be positive roots in a crystallographic root system \mathfrak{R}, such that their sums yield an equality

i=0mβi=γ0+γ1\sum_{i=0}^{m}\beta_{i}=\gamma_{0}+\gamma_{1}

between two positive roots. Then for some 0jm0\leq j\leq m the smaller sum i=0,ijmβi\sum_{i=0,i\neq j}^{m}\beta_{i} is also a positive root (or zero if m=0m=0) and for some t{0,1}t\in\{0,1\} either

i=0,ijmβiγtorβjγt\sum_{i=0,i\neq j}^{m}\beta_{i}-\gamma_{t}\qquad\textrm{or}\qquad\beta_{j}-\gamma_{t}

lies in +{0}\mathfrak{R}_{+}\sqcup\{0\}.

Proof.

This now follows by combining the previous proposition with Lemma 2.2(i). ∎

This is close to what we need; the next lemma refines this statement to show that if β0\beta_{0} is a simple root, then we can ensure that this root does not appear in the conclusion.

Lemma 2.9.

Suppose that α\alpha is a simple root and β1,,βm\beta_{1},\ldots,\beta_{m} for m1m\geq 1 and γ\gamma are positive roots in a crystallographic root system, such that

α+i=1mβiandα+i=1mβiγare both positive roots.\alpha+\sum_{i=1}^{m}\beta_{i}\qquad\textrm{and}\qquad\alpha+\sum_{i=1}^{m}\beta_{i}-\gamma\qquad\textrm{are both positive roots}.

Furthermore suppose that for any subset {i1,,ik}\{i_{1},\ldots,i_{k}\} of {1,,m}\{1,\ldots,m\}, the expression

α+j=1kβijγis neither a negative root nor zero.\alpha+\sum_{j=1}^{k}\beta_{i_{j}}-\gamma\qquad\textrm{is neither a negative root nor zero}. (2.3)

Then there exists a subset {i1,,ik}\{i_{1},\ldots,i_{k}\} of {1,,m}\{1,\ldots,m\} with the property that

j=1kβijγis a positive root or is zero.\sum_{j=1}^{k}\beta_{i_{j}}-\gamma\qquad\textrm{is a positive root or is zero}.
Proof.

We induct on mm, so we presume that the claim is true <m<m. Let 0km0\leq k\leq m be the largest integer such that both γj=1kβij\gamma-\sum_{j=1}^{k}\beta_{i_{j}} and α+j=k+1mβij\alpha+\sum_{j=k+1}^{m}\beta_{i_{j}} are positive roots, for some partition

{i1,,ik},{ik+1,,im}\{i_{1},\ldots,i_{k}\},\qquad\{i_{k+1},\ldots,i_{m}\}

of {1,,m}\{1,\ldots,m\} into two subsets. If k=mk=m then γj=1mβij+\gamma-\sum_{j=1}^{m}\beta_{i_{j}}\in\mathfrak{R}_{+}, but this combines with the initial assumption α+i=1mβiγ+\alpha+\sum_{i=1}^{m}\beta_{i}-\gamma\in\mathfrak{R}_{+} and simpleness of α\alpha to a contradiction, so k<mk<m. By Lemma 2.2(iii) there then exists an integer k+1tmk+1\leq t\leq m such that

α+j=k+1,jtmβijis a positive root.\alpha+\sum_{j=k+1,j\neq t}^{m}\beta_{i_{j}}\qquad\textrm{is a positive root}. (2.4)

Now applying Proposition 2.7 to the equality of roots

(α+j=k+1,jtmβij)+βit=(α+i=jmβjγ)+(γj=1kβij)(\alpha+\sum_{j=k+1,j\neq t}^{m}\beta_{i_{j}})+\beta_{i_{t}}=(\alpha+\sum_{i=j}^{m}\beta_{j}-\gamma)+(\gamma-\sum_{j=1}^{k}\beta_{i_{j}})

yields that at least one of the expressions

γβitj=1kβij,βit+j=1kβijγ,α+j=1,jtmβijγ,γαj=1,jtmβij,\gamma-\beta_{i_{t}}-\sum_{j=1}^{k}\beta_{i_{j}},\qquad\beta_{i_{t}}+\sum_{j=1}^{k}\beta_{i_{j}}-\gamma,\qquad\alpha+\sum_{j=1,j\neq t}^{m}\beta_{i_{j}}-\gamma,\qquad\gamma-\alpha-\sum_{j=1,j\neq t}^{m}\beta_{i_{j}},

lies in +{0}\mathfrak{R}_{+}\sqcup\{0\}. In the first case, if this expression equals zero we’re done and if it’s a positive root then combining this with equation (2.4) implies that kk was not maximal. The second case would yield the claim immediately. For the third case, the assumption of (2.3) implies that this expression can’t be zero and then the claim would follow from the induction hypothesis on mm. The fourth case is excluded by the same assumption. ∎

Corollary 2.10.

Suppose that α\alpha is a simple root and β1,,βm\beta_{1},\ldots,\beta_{m} for m1m\geq 1 and γ0,γ1\gamma_{0},\gamma_{1} are positive roots in a crystallographic root system, such that there is an equality

α+i=1mβi=γ0+γ1\alpha+\sum_{i=1}^{m}\beta_{i}=\gamma_{0}+\gamma_{1}

of positive roots. Then for some subset {i1,,ik}\{i_{1},\ldots,i_{k}\} of {1,,m}\{1,\ldots,m\} and some t{0,1}t\in\{0,1\}, the expression

j=1kβijγt\sum_{j=1}^{k}\beta_{i_{j}}-\gamma_{t}

is a positive root or is zero.

Proof.

We may assume that γtα\gamma_{t}\neq\alpha for both t{0,1}t\in\{0,1\} (otherwise the claim is immediate), and we first invoke Corollary 2.8 with β0=α\beta_{0}=\alpha. If in its conclusion j=0j=0, then the claim follows immediately as αγt\alpha-\gamma_{t} in +{0}\mathfrak{R}_{+}\sqcup\{0\} would imply that γt=α\gamma_{t}=\alpha. Suppose on the other hand that j0j\neq 0 and that moreover βjγt\beta_{j}-\gamma_{t} is not in +{0}\mathfrak{R}_{+}\sqcup\{0\} (otherwise the claim follows). Then this corollary states that for certain γt\gamma_{t}, the expression

α+i=1,ijmβiγtlies in +{0}.\alpha+\sum_{i=1,i\neq j}^{m}\beta_{i}-\gamma_{t}\qquad\textrm{lies in }\qquad\mathfrak{R}_{+}\sqcup\{0\}.

If this expression is zero then βj=γt\beta_{j}=\gamma_{t^{\prime}}, where t:=1tt^{\prime}:=1-t, and the claim follows for γt\gamma_{t^{\prime}} so we may assume that this expression is a positive root. In fact, from the equation

k+1mβijγt=(α+i=1kβijγt)\sum_{k+1}^{m}\beta_{i_{j}}-\gamma_{t^{\prime}}=-\bigl{(}\alpha+\sum_{i=1}^{k}\beta_{i_{j}}-\gamma_{t}\bigr{)}

it follows may furthermore assume that condition (2.3) holds, as again the claim would otherwise follow for γt\gamma_{t^{\prime}}; then we may invoke the previous lemma, which yields the claim. ∎

Notation 2.11.

Given two positive roots γ\gamma and γ\gamma^{\prime} in some root system, we write γ<γ\gamma^{\prime}<\gamma if their difference γγ\gamma-\gamma^{\prime} lies in the convex cone of positive roots.

Finally, we prove part (i) of the main Lemma:

Proof.

As the simple root α\alpha is not in w\mathfrak{R}_{w} by assumption, the root w(α)w(\alpha) is positive. Given an integer m0m\geq 0 and some roots β1,,βmw\beta_{1},\ldots,\beta_{m}\in\mathfrak{R}_{w} such that w(α+i=1mβi)w(\alpha+\sum_{i=1}^{m}\beta_{i}) is a positive root but is not simple, we will find roots β1,,βmw\beta_{1}^{\prime},\ldots,\beta_{m^{\prime}}^{\prime}\in\mathfrak{R}_{w} (with mm+1m^{\prime}\leq m+1) such that w(α+i=1mβi)w(\alpha+\sum_{i=1}^{m^{\prime}}\beta_{i}^{\prime}) is still a positive root, but is smaller in the sense that

w(α+i=1mβi)<w(α+i=1mβi)=:γ~.w(\alpha+\sum_{i=1}^{m^{\prime}}\beta_{i}^{\prime})<w(\alpha+\sum_{i=1}^{m}\beta_{i})=:\tilde{\gamma}.

By downwards induction on height, the claim then follows. Since γ~\tilde{\gamma} is not simple, we may split γ~=γ~0+γ~1\tilde{\gamma}=\tilde{\gamma}_{0}+\tilde{\gamma}_{1} into some other positive roots γ~0,γ~1\tilde{\gamma}_{0},\tilde{\gamma}_{1}.

If both roots w1(γ~0)w^{-1}(\tilde{\gamma}_{0}) and w1(γ~1)w^{-1}(\tilde{\gamma}_{1}) are positive then by the previous corollary (setting γi:=w1(γ~i)\gamma_{i}:=w^{-1}(\tilde{\gamma}_{i})), for at least one t{0,1}t\in\{0,1\} there is a subset {i1,,ik}\{i_{1},\ldots,i_{k}\} of {1,,m}\{1,\ldots,m\} such that i=1kβijw1(γ~t)\sum_{i=1}^{k}\beta_{i_{j}}-w^{-1}(\tilde{\gamma}_{t}) is a positive root or is zero; in the former case, it moreover lies in w\mathfrak{R}_{w} as

w(i=1kβijw1(γ~t))=i=1kw(βij)γ~tw\bigl{(}\sum_{i=1}^{k}\beta_{i_{j}}-w^{-1}(\tilde{\gamma}_{t})\bigr{)}=\sum_{i=1}^{k}w(\beta_{i_{j}})-\tilde{\gamma}_{t}

is a sum of negative roots.

If on the other hand some w1(γ~t)w^{-1}(\tilde{\gamma}_{t}) is negative, then w1(γ~t)-w^{-1}(\tilde{\gamma}_{t}) lies in w\mathfrak{R}_{w} so the same conclusion holds with k=0k=0. Hence in either case

w(α+i=k+1mβij+(i=1kβijw1(γ~t)))=w(α+i=1mβiw1(γ~t))=γ~t<γ~,w\Bigl{(}\alpha+\sum_{i=k+1}^{m}\beta_{i_{j}}+\bigl{(}\sum_{i=1}^{k}\beta_{i_{j}}-w^{-1}(\tilde{\gamma}_{t})\bigr{)}\Bigr{)}=w\bigl{(}\alpha+\sum_{i=1}^{m}\beta_{i}-w^{-1}(\tilde{\gamma}_{t})\bigr{)}=\tilde{\gamma}_{t^{\prime}}<\tilde{\gamma},

where t:=1tt^{\prime}:=1-t. ∎

Example 2.12.

Consider w=s1s2w=s_{1}s_{2} in type 𝖡2\mathsf{B}_{2}. Then α1w\alpha_{1}\notin\mathfrak{R}_{w} and as crossw(α1)={α122,α2}\mathrm{cross}_{w}(\alpha_{1})=\{\alpha_{122},\alpha_{2}\} it follows that there is no “path of simple roots” within crossw(α1)\mathrm{cross}_{w}(\alpha_{1}) from w(α1)=α122w(\alpha_{1})=\alpha_{122} down to a simple root.

Remark 2.13.

I am not sure whether these results naturally generalise to the noncrystallographic case; consider for any positive root β\beta the set

crossw(β):={w(c0β+i=1mciβi):c0,,cm>0,β1,,βmw,m0}+,\mathrm{cross}_{w}(\beta):=\bigl{\{}w(c_{0}\beta+\sum_{i=1}^{m}c_{i}\beta_{i})\in\mathfrak{R}:c_{0},\ldots,c_{m}\in\mathbb{R}_{>0},\beta_{1},\ldots,\beta_{m}\in\mathfrak{R}_{w},m\geq 0\bigr{\}}\cap\mathfrak{R}_{+},

where one might want to put some restrictions on the cic_{i}. For 𝖧3\mathsf{H}_{3} and 𝖧4\mathsf{H}_{4} there is a well-known “folding” argument showing that one can realise their root systems inside those of 𝖣6\mathsf{D}_{6} and 𝖤8\mathsf{E}_{8}, preserving simple roots and embedding the corresponding reflection groups; part (i) of the main Lemma then holds for c0=1c_{0}=1 and ci{1,φ±1}c_{i}\in\{1,\varphi^{\pm 1}\} for i>0i>0, where φ\varphi denotes the golden ratio. On the other hand, if we fix c0=1c_{0}=1 then part (i) of the main Lemma fails for say Coxeter elements of minimal length in type 𝖨2(7)\mathsf{I}_{2}(7). For dihedral groups there are only two simple roots and in nontrivial cases the other simple root lies in w\mathfrak{R}_{w}; thus for (i) a positive linear combination of the simple roots can be used if there are no restrictions on c0c_{0}. However, the main point of (i) was to obtain (ii) by combining it with braid invariance, but braid invariance still fails in 𝖧3\mathsf{H}_{3}.

Example 2.14.

Consider w=s2s1w=s_{2}s_{1} in type 𝖧3\mathsf{H}_{3} and denote the golden ratio by φ\varphi. Then w={α1,φα1+α2}\mathfrak{R}_{w}=\{\alpha_{1},\varphi\alpha_{1}+\alpha_{2}\} and the only simple root in crossw(α2)\mathrm{cross}_{w}(\alpha_{2}) is

α1=w(φα1+φα2)=w(α2+φ1α1+φ1(φα1+α2)).\alpha_{1}=w(\varphi\alpha_{1}+\varphi\alpha_{2})=w\bigl{(}\alpha_{2}+\varphi^{-1}\alpha_{1}+\varphi^{-1}(\varphi\alpha_{1}+\alpha_{2})\bigr{)}.

2.3 Crossing for the Deligne-Garside normal form

In this subsection we prove the remainder of the main Lemma, Corollary A and the main Proposition.

Recall (e.g. from [Mal21, §4.1]) that two elements w2,w1w_{2},w_{1} of a twisted Weyl group are in right Deligne-Garside normal form (modulo moving twists) if and only if for any simple reflection sis_{i} with (w2si)<(w2)\ell(w_{2}s_{i})<\ell(w_{2}) we also have (siw1)<(w1)\ell(s_{i}w_{1})<\ell(w_{1}). Furthermore, we will repeatedly use the following property from [Mal21, Proposition C(ii)]: for any element ww of a twisted Weyl group and integer d0d\geq 0, we have

DG(bwd)stw=.\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}\cap\mathfrak{R}_{\mathrm{st}}^{w}=\varnothing. (2.5)

From part (i) of the main Lemma we deduce

Corollary 2.15.

Let wm,,w1w_{m},\ldots,w_{1} be elements of a twisted Weyl group WW such that their product bwmbw1b_{w_{m}}\cdots b_{w_{1}} is in Deligne-Garside normal form, after moving twists. Then for any yy in WW such that w1yw_{1}\geq y and any simple root α\alpha such that β:=y1(α)\beta:=y^{-1}(\alpha) lies in +\w1\mathfrak{R}_{+}\backslash\mathfrak{R}_{w_{1}}, the set

crosswm1crossw1(β)\mathrm{cross}_{w_{m-1}}\cdots\mathrm{cross}_{w_{1}}(\beta)

contains simple roots not lying in wm\mathfrak{R}_{w_{m}}.

Proof.

By induction on mm, it suffices to show that crossw1(β)\mathrm{cross}_{w_{1}}(\beta) contains a simple root not in w2\mathfrak{R}_{w_{2}}. By assumption we have a reduced decomposition w1=xyw_{1}=xy, and the root w1(β)=x(α)w_{1}(\beta)=x(\alpha) is positive so α\alpha is not in x\mathfrak{R}_{x}. By part (i) of the main Lemma there then exists a simple root αcrossx(α)\alpha^{\prime}\in\mathrm{cross}_{x}(\alpha), which means that

x(α+i=1mβi)=αx(\alpha+\sum_{i=1}^{m}\beta_{i})=\alpha^{\prime}

for some β1,,βmx\beta_{1},\ldots,\beta_{m}\in\mathfrak{R}_{x}. As x=w1y1x=w_{1}y^{-1}, this yields the equation

w1(β+i=1my1(βi))=αw_{1}\bigr{(}\beta+\sum_{i=1}^{m}y^{-1}(\beta_{i})\bigl{)}=\alpha^{\prime}

with each y1(βi)y1(x)w1y^{-1}(\beta_{i})\in y^{-1}(\mathfrak{R}_{x})\subseteq\mathfrak{R}_{w_{1}} which then implies that αcrossw1(β)\alpha^{\prime}\in\mathrm{cross}_{w_{1}}(\beta). The normal form condition on the pair w2,w1w_{2},w_{1} is equivalent to requiring that w11(α~)w_{1}^{-1}(\tilde{\alpha}) is a negative root for any simple root α~\tilde{\alpha} lying in w2\mathfrak{R}_{w_{2}}. As w11(α)=β+i=1my1(βi)w_{1}^{-1}(\alpha^{\prime})=\beta+\sum_{i=1}^{m}y^{-1}(\beta_{i}) is a sum of positive roots and is therefore positive, it now follows that α\alpha^{\prime} does not lie in w2\mathfrak{R}_{w_{2}}. ∎

When α\alpha is not simple, there may not be simple roots in these sets:

Example 2.16.

Consider again w=s1s2s3s1s2w=s_{1}s_{2}s_{3}s_{1}s_{2} in type 𝖡3\mathsf{B}_{3}. Then DGN(bwi)=bwi\mathrm{DGN}(b_{w}^{i})=b_{w}^{i} and the root α1233\alpha_{1233} does not lie in ww\mathfrak{R}_{w}\sqcup\mathfrak{R}^{w}, yet

crossw(α1233)={w(α1233)}={α233}w.\mathrm{cross}_{w}(\alpha_{1233})=\{w(\alpha_{1233})\}=\{\alpha_{233}\}\in\mathfrak{R}_{w}.
Notation 2.17.

Recall that we write DG>1(bwd):=bwdDG(bwd)1\mathrm{DG}_{>1}(b_{w}^{d}):=b_{w}^{d}\mathrm{DG}(b_{w}^{d})^{-1} for the left complement to DG(bwd)\mathrm{DG}(b_{w}^{d}) in bwdb_{w}^{d}.

Corollary 2.18.

Let ww be an element of a twisted Weyl group WW, let 𝔑+\mathfrak{N}\subseteq\mathfrak{R}_{+} be a subset of positive roots and pick an integer d0d\geq 0.

  1. (i)

    If 𝔑DG(bwd)\mathfrak{N}\subseteq\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})} then crosswd(𝔑)=\mathrm{cross}_{w}^{d}(\mathfrak{N})=\varnothing.

  2. (ii)

    If there exist a simple root α\alpha in 𝔑\mathfrak{N} and element yy in WW satisfying

    DG(bwd)yandy1(α)+\DG(bwd),\mathrm{DG}(b_{w}^{d})\geq y\qquad\textrm{and}\qquad y^{-1}(\alpha)\in\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})},

    then crosswd(𝔑)\mathrm{cross}_{w}^{d}(\mathfrak{N}) contains simple roots.

Proof.

(i): If 𝔑DG(bwd)\mathfrak{N}\subseteq\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})} then by Lemma 2.6 we have

crosswd(𝔑)=crossDG>1(bwd)crossDG(bwd)(𝔑)=.\mathrm{cross}_{w}^{d}(\mathfrak{N})=\mathrm{cross}_{\mathrm{DG}_{>1}(b_{w}^{d})}\mathrm{cross}_{\mathrm{DG}(b_{w}^{d})}(\mathfrak{N})=\varnothing.

(ii): Corollary 2.15 yields that there are simple roots lying in

crossDG>1(bwd)crossDG(bwd)(𝔑)=crosswd(𝔑).\mathrm{cross}_{\mathrm{DG}_{>1}(b_{w}^{d})}\mathrm{cross}_{\mathrm{DG}(b_{w}^{d})}(\mathfrak{N})=\mathrm{cross}_{w}^{d}(\mathfrak{N}).\qed

From this corollary we can deduce part (ii) of the main Lemma:

Proof.

The implication \Rightarrow follows immediately from the first part of the corollary. For \Leftarrow, suppose that the inequality DG(bwd)w\mathrm{DG}(b_{w}^{d})\geq w^{\prime} does not hold. So we may suppose that there exists a simple reflection sis_{i} and an element yWy\in W such that

DG(bwd)y,wsiy,DG(bwd)siy.\mathrm{DG}(b_{w}^{d})\geq y,\qquad w^{\prime}\geq s_{i}y,\qquad\mathrm{DG}(b_{w}^{d})\not\geq s_{i}y.

Then y1(αi)y^{-1}(\alpha_{i}) is a positive root lying in w\DG(bwd)\mathfrak{R}_{w^{\prime}}\backslash\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}, so from the second part of the corollary it follows that crosswd(w)\mathrm{cross}_{w}^{d}(\mathfrak{R}_{w^{\prime}}) contains a simple root. ∎

Lemma 2.19.

Let ww be an element of a twisted Weyl group and pick an integer d0d\geq 0. Then for any simple root α\alpha whose orbit under ww consists solely of other simple roots, the root DG(bwd)(α)\mathrm{DG}(b_{w}^{d})(\alpha) is again simple.

Proof.

From equation (2.5) it follows that α\alpha is not in DG(bwd)\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}, so part (i) of the main Lemma implies that there has to exist a simple root α′′\alpha^{\prime\prime} in crossDG(bwd)(α)\mathrm{cross}_{\mathrm{DG}(b_{w}^{d})}(\alpha), which means that there exist roots β1,1,,βm,1DG(bwd)\beta_{1,1},\ldots,\beta_{m,1}\in\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})} such that

DG(bwd)(α+i=1mβi,1)=α′′.\mathrm{DG}(b_{w}^{d})(\alpha+\sum_{i=1}^{m}\beta_{i,1})=\alpha^{\prime\prime}.

In particular, if DG(bwd)(α)\mathrm{DG}(b_{w}^{d})(\alpha) is not simple then m>0m>0. From Corollary 2.15 we similarly obtain a sequence in crossDG>1(bwd)(α,α′′)\mathrm{cross}_{\mathrm{DG}_{>1}(b_{w}^{d})}(\alpha^{\prime},\alpha^{\prime\prime}) for some simple root α\alpha^{\prime}, which we may concatenate with the (βi,1)(\beta_{i,1}) to a sequence (βi,j)crossDGN(bwd)(α,α)(\beta_{i,j})\in\mathrm{cross}_{\mathrm{DGN}(b_{w}^{d})}(\alpha^{\prime},\alpha). Since the (βi,1)(\beta_{i,1}) part of this sequence is nontrivial, Lemma 2.6 implies that we can transfer (βi,j)(\beta_{i,j}) to a nontrivial sequence (βi,j)(\beta_{i,j}^{\prime}) in crosswd(α,α)\mathrm{cross}_{w}^{d}(\alpha^{\prime},\alpha). This gives positive roots αj\alpha_{j}^{\prime} inductively defined for 1jd1\leq j\leq d as

αj:=w(αj1+iβi,j),α0:=α,\alpha_{j}^{\prime}:=w(\alpha_{j-1}^{\prime}+\sum_{i}\beta_{i,j}^{\prime}),\qquad\alpha_{0}^{\prime}:=\alpha,

with each βi,jw\beta_{i,j}^{\prime}\in\mathfrak{R}_{w}. A priori the roots αj\alpha_{j}^{\prime} are not necessarily simple, but we now inductively prove that they are all simple roots lying in the ww-orbit of α\alpha, and that the elements βi,j\beta_{i,j}^{\prime} are all zero: if it’s true <j<j then the induction hypothesis yields that

αj=w(αj1)+iw(βi,j)+\alpha_{j}^{\prime}=w(\alpha_{j-1}^{\prime})+\sum_{i}w(\beta_{i,j}^{\prime})\in\mathfrak{R}_{+}

but as w(αj1)w(\alpha_{j-1}^{\prime}) is simple by assumption and each nontrivial root βi,j\beta_{i,j}^{\prime} lies in w\mathfrak{R}_{w} this implies that each βi,j=0\beta_{i,j}^{\prime}=0 and then αj=w(αj1)\alpha^{\prime}_{j}=w(\alpha_{j-1}^{\prime}). But that is a contradiction as the sequence (βi,j)(\beta_{i,j}^{\prime}) was constructed to be nontrivial, and therefore DG(bwd)(α)\mathrm{DG}(b_{w}^{d})(\alpha) must be a simple root. ∎

Corollary 2.20.

If ww is convex (resp. firmly convex), then each of the elements in the sequence

w,DG(bw2)wDG(bw2)1,DG(bw3)wDG(bw3)1,,w,\qquad\mathrm{DG}(b_{w}^{2})\,w\,\mathrm{DG}(b_{w}^{2})^{-1},\qquad\mathrm{DG}(b_{w}^{3})\,w\,\mathrm{DG}(b_{w}^{3})^{-1},\qquad\ldots, (2.6)

of cyclic shifts is convex (resp. firmly convex).

Proof.

It was proven in [Mal21, Proposition 4.39] that conjugation by DG(bwd)\mathrm{DG}(b_{w}^{d}) induces a sequence of cyclic shifts, so from [Mal21, Proposition A(i)] we then deduce that

stDG(bwd)wDG(bwd)1=DG(bwd)(stw).\mathfrak{R}_{\mathrm{st}}^{\mathrm{DG}(b_{w}^{d})w\mathrm{DG}(b_{w}^{d})^{-1}}=\mathrm{DG}(b_{w}^{d})(\mathfrak{R}_{\mathrm{st}}^{w}).

If ww is convex then by height considerations it follows that it must map any of the simple roots in stw+\mathfrak{R}_{\mathrm{st}}^{w}\cap\mathfrak{R}_{+} to other simple roots. The lemma now implies that DG(bwd)(stw)\mathrm{DG}(b_{w}^{d})(\mathfrak{R}_{\mathrm{st}}^{w}) is also a standard parabolic subsystem. ∎

The main Proposition follows from

Proposition 2.21.

Let ww be an element of a twisted Weyl group.

  1. (i)

    Let 𝔑=y\mathfrak{N}=\mathfrak{R}_{y} for some yWy\in W. Then 𝔑\mathfrak{N} is ww-nimble if and only if ywy\geq w and yyw1y\geq yw^{-1} in the weak left Bruhat-Chevalley order.

    In particular, the inversion sets associated to the elements in the sequence

    w=DG(bw),DG(bw2),DG(bw3),,w=\mathrm{DG}(b_{w}),\qquad\mathrm{DG}(b_{w}^{2}),\qquad\mathrm{DG}(b_{w}^{3}),\qquad\ldots, (2.7)

    yield ww-nimble sets. On the other hand, +\stw\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w} is ww-nimble if and only if ww is convex.

  2. (ii)

    If indeed ww is convex, then

    +\stw+stw+\stwandDG(bwd)+stwDG(bwd)\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}+\mathfrak{R}_{\mathrm{st}}^{w}\subseteq\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}\qquad\textrm{and}\qquad\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}+\mathfrak{R}_{\mathrm{st}}^{w}\subseteq\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}

    for any natural number d1d\geq 1.

Proof.

(i): The inclusion wy\mathfrak{R}_{w}\subseteq\mathfrak{R}_{y} is equivalent to ywy\geq w. Under this assumption, there is a reduced decomposition y=(yw1)wy=(yw^{-1})w which means

y=w1(yw1)w\mathfrak{R}_{y}=w^{-1}(\mathfrak{R}_{yw^{-1}})\sqcup\mathfrak{R}_{w}

and applying ww to this identity then yields

w(y)+=yw1(w(w)+)=yw1,w(\mathfrak{R}_{y})\cap\mathfrak{R}_{+}=\mathfrak{R}_{yw^{-1}}\sqcup\bigl{(}w(\mathfrak{R}_{w})\cap\mathfrak{R}_{+}\bigr{)}=\mathfrak{R}_{yw^{-1}},

which reduces the nimbleness condition to the inclusion yw1=w(y)+y\mathfrak{R}_{yw^{-1}}=w(\mathfrak{R}_{y})\cap\mathfrak{R}_{+}\subseteq\mathfrak{R}_{y}; this yields the first two claims.

From wstw=\mathfrak{R}_{w}\cap\mathfrak{R}_{\mathrm{st}}^{w}=\varnothing and w(stw)=stww(\mathfrak{R}_{\mathrm{st}}^{w})=\mathfrak{R}_{\mathrm{st}}^{w} we obtain the inclusions

w+\stwandw(+\stw)++\stw.\mathfrak{R}_{w}\subseteq\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}\qquad\textrm{and}\qquad w(\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w})\cap\mathfrak{R}_{+}\subseteq\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}.

Thus +\stw\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w} is ww-nimble if and only if +\stw\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w} is convex, so the final claim follows from [Mal21, Proposition 4.21].

(ii): The first identity follows from the assumption that stw\mathfrak{R}_{\mathrm{st}}^{w} forms a standard parabolic subsystem. For the final one, let βDG(bwd)\beta\in\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})} and γstw\gamma\in\mathfrak{R}_{\mathrm{st}}^{w} and suppose that c0β+c1γc_{0}\beta+c_{1}\gamma is a root for some c0,c1>0c_{0},c_{1}\in\mathbb{R}_{>0}. As stw\mathfrak{R}_{\mathrm{st}}^{w} is a standard parabolic subroot system and β\beta is positive, equation (2.5) implies that c0β+c1γc_{0}\beta+c_{1}\gamma must be a positive root. If γ\gamma is negative then the same equation implies that DG(bwd)(γ)\mathrm{DG}(b_{w}^{d})(\gamma) is still negative and therefore so is

DG(bwd)(c0β+c1γ)=c0DG(bwd)(β)+c1DG(bwd)(γ).\mathrm{DG}(b_{w}^{d})(c_{0}\beta+c_{1}\gamma)=c_{0}\mathrm{DG}(b_{w}^{d})(\beta)+c_{1}\mathrm{DG}(b_{w}^{d})(\gamma).

On the other hand, if γ\gamma is positive then the previous lemma implies that DG(bwd)(γ)\mathrm{DG}(b_{w}^{d})(\gamma) lies in the positive half of the standard parabolic subsystem DG(bwd)(stw)\mathrm{DG}(b_{w}^{d})(\mathfrak{R}_{\mathrm{st}}^{w}). If furthermore DG(bwd)(c0β+c1γ)\mathrm{DG}(b_{w}^{d})(c_{0}\beta+c_{1}\gamma) is positive then as DG(bwd)(β)\mathrm{DG}(b_{w}^{d})(\beta) is negative it must lie in the negative half of DG(bwd)(stw)\mathrm{DG}(b_{w}^{d})(\mathfrak{R}_{\mathrm{st}}^{w}), but then β\beta lies in stw\mathfrak{R}_{\mathrm{st}}^{w} which contradicts equation (2.5) again. ∎

By [Mal21, Proposition C(i)], the sequence (2.7) (and hence also (2.6)) stabilises after |+\w|(w)|\mathfrak{R}_{+}\backslash\mathfrak{R}^{w}|-\ell(w) terms; we will reprove this at the end of this subsection.

Example 2.22.

Consider w=s3s2s1w=s_{3}s_{2}s_{1} and w=s2ww^{\prime}=s_{2}w in type 𝖠3\mathsf{A}_{3}. Then ww is elliptic so it is convex and

w=DG(bw3)>w>w,w_{\circ}=\mathrm{DG}(b_{w}^{3})>w^{\prime}>w,

but as w(α3)=α2w(\alpha_{3})=\alpha_{2} we have

w(w)+ww(\mathfrak{R}_{w^{\prime}})\cap\mathfrak{R}_{+}\not\subseteq\mathfrak{R}_{w^{\prime}}

so w\mathfrak{R}_{w^{\prime}} is not nimble.

Typically however, the sets w(w+)\mathfrak{R}_{w}\sqcup(\mathfrak{R}^{w}\cap\mathfrak{R}_{+}) and w(stw+)\mathfrak{R}_{w}\sqcup(\mathfrak{R}_{\mathrm{st}}^{w}\cap\mathfrak{R}_{+}) are not convex:

Example 2.23.

Consider s2s_{2} in type 𝖡2\mathsf{B}_{2}. It is not firmly convex, and

w={α2},w+={α12},w+(w+)={α122}.\mathfrak{R}_{w}=\{\alpha_{2}\},\qquad\mathfrak{R}^{w}\cap\mathfrak{R}_{+}=\{\alpha_{12}\},\qquad\mathfrak{R}_{w}+(\mathfrak{R}^{w}\cap\mathfrak{R}_{+})=\{\alpha_{122}\}.
Example 2.24.

Consider s3s1s2s1s_{3}s_{1}s_{2}s_{1} in type 𝖡3\mathsf{B}_{3}. It is not firmly convex, and

w={α23},stw+={α1,α12,α2,α123},w+(stw+)={α12233}.\mathfrak{R}_{w}=\{\alpha_{23}\},\qquad\mathfrak{R}_{\mathrm{st}}^{w}\cap\mathfrak{R}_{+}=\{\alpha_{1},\alpha_{12},\alpha_{2},\alpha_{123}\},\qquad\mathfrak{R}_{w}+(\mathfrak{R}_{\mathrm{st}}^{w}\cap\mathfrak{R}_{+})=\{\alpha_{12233}\}.
Lemma 2.25.

Let ww be an element of a twisted Weyl group and pick a natural number d0d\geq 0. Then the following are equivalent:

  1. (i)

    The set +\(stwDG(bwd))\mathfrak{R}_{+}\backslash(\mathfrak{R}_{\mathrm{st}}^{w}\sqcup\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}) is empty,

  2. (ii)

    The set +\(stwDG(bwd))\mathfrak{R}_{+}\backslash(\mathfrak{R}_{\mathrm{st}}^{w}\sqcup\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}) does not contain any simple roots,

  3. (iii)

    The set crosswd(+\stw)\mathrm{cross}_{w}^{d}(\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}) is empty,

  4. (iv)

    The set crosswd(+\stw)\mathrm{cross}_{w}^{d}(\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}) does not contain any simple roots.

Proof.

(ii) \Rightarrow (i): This follows from combining [Mal21, Lemma 3.19(i)] with Lemma 2.19.

(iv) \Rightarrow (ii): If +\(stwDG(bwd))\mathfrak{R}_{+}\backslash(\mathfrak{R}_{\mathrm{st}}^{w}\sqcup\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}) contains simple roots then by Lemma 2.6 and Corollary 2.15 so does

crosswd(+\stw)=crossDG>1(bwd)crossDG(bwd)(+\stw).\mathrm{cross}_{w}^{d}(\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w})=\mathrm{cross}_{\mathrm{DG}_{>1}(b_{w}^{d})}\mathrm{cross}_{\mathrm{DG}(b_{w}^{d})}(\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}).

(i) \Rightarrow (iii) follows from the same identity, and (iii) \Rightarrow (iv) is immediate. ∎

Example 2.26.

Let ww be reflection in a non-simple root in type 𝖡2\mathsf{B}_{2}, then w(w+)=+\mathfrak{R}_{w}\sqcup(\mathfrak{R}^{w}\cap\mathfrak{R}_{+})=\mathfrak{R}_{+}.

Example 2.27.

Consider the elements w=s2s3s2s1w=s_{2}s_{3}s_{2}s_{1} and w=s2s1ww^{\prime}=s_{2}s_{1}w in type 𝖡3\mathsf{B}_{3}. Then

ws3=DG(bw2)>w>w,w_{\circ}s_{3}=\mathrm{DG}(b_{w}^{2})>w^{\prime}>w,

but +\(ww)={α23,α233}\mathfrak{R}_{+}\backslash(\mathfrak{R}_{w^{\prime}}\sqcup\mathfrak{R}^{w})=\{\alpha_{23},\alpha_{233}\}.

We deduce part (iii) of the main Lemma:

Proof.

If ww is convex then the claim follows from part (ii) of the main Lemma, combined with the fact that pb(w)\mathrm{pb}(w) is an upper bound for DG(bwd)\mathrm{DG}(b_{w}^{d}). If on the other hand it is not convex, then +\stw\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w} is not convex by [Mal21, Proposition 4.21] again so +\(stwDG(bwd))\mathfrak{R}_{+}\backslash(\mathfrak{R}_{\mathrm{st}}^{w}\sqcup\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}) must be nonempty (as DG(bwd)\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})} is convex), and then the claim follows from the previous lemma. ∎

Lemma 2.28.

Let ww be an element of a twisted Weyl group and let 𝔑\mathfrak{N} a ww-nimble set also containing w1\mathfrak{R}_{w^{-1}}. Then 𝔑\mathfrak{N} is also nimble for the element w1w^{-1}, and for any integer d0d\geq 0 we have

crosswd(𝔑)=if and only if crossw1d(𝔑)=.\mathrm{cross}_{w}^{d}(\mathfrak{N})=\varnothing\qquad\textrm{if and only if }\qquad\mathrm{cross}_{w^{-1}}^{d}(\mathfrak{N})=\varnothing.
Proof.

Pick a root β\beta in 𝔑\w1\mathfrak{N}\backslash\mathfrak{R}_{w^{-1}}. If β\beta also lies in stw1=stw\mathfrak{R}_{\mathrm{st}}^{w^{-1}}=\mathfrak{R}_{\mathrm{st}}^{w}, then by nimbleness of ww and positivity the root w1(β)=word(w)1(β)w^{-1}(\beta)=w^{\mathrm{ord}(w)-1}(\beta) also lies in 𝔑\mathfrak{N}. If on the other hand it does not lie in stw1\mathfrak{R}_{\mathrm{st}}^{w^{-1}}, then wi(β)w1𝔑w^{-i}(\beta)\in\mathfrak{R}_{w^{-1}}\subseteq\mathfrak{N} for some i>0i>0. Applying wi1w^{i-1}, nimbleness yields that w1(β)w^{-1}(\beta) also lies in 𝔑\mathfrak{N}.

If say crosswd(𝔑)\mathrm{cross}_{w}^{d}(\mathfrak{N})\neq\varnothing then there exists a sequence of roots (βi,j)crosswd(β,β)(\beta_{i,j})\in\mathrm{cross}_{w}^{d}(\beta^{\prime},\beta) for some roots β,β\beta,\beta^{\prime} in 𝔑\mathfrak{N}. As w(w)=w1-w(\mathfrak{R}_{w})=\mathfrak{R}_{w^{-1}} this yields another sequence

w(βd+1i,j)crossw1d(β,β),-w(\beta_{d+1-i,j})\in\mathrm{cross}_{w^{-1}}^{d}(\beta,\beta^{\prime}),

implying that crossw1d(𝔑)\mathrm{cross}_{w^{-1}}^{d}(\mathfrak{N})\neq\varnothing. ∎

Example 2.29.

Consider w=s3s2s3s1w=s_{3}s_{2}s_{3}s_{1} in type 𝖡3\mathsf{B}_{3}. It does not fix any roots, and the set

𝔑:=ww1=+\{α2,α12,α12233}\mathfrak{N}:=\mathfrak{R}_{w}\cup\mathfrak{R}_{w^{-1}}=\mathfrak{R}_{+}\backslash\{\alpha_{2},\alpha_{12},\alpha_{12233}\}

is nimble.

Corollary A follows from

Corollary 2.30.

Let ww be an element of a twisted Weyl group WW and let yy be an element of WW such that yw1y\geq w^{-1}. Then for any integer d0d\geq 0 we have

DG(bwd)yif and only if DG(bw1d)y.\mathrm{DG}(b_{w}^{d})\geq y\qquad\textrm{if and only if }\qquad\mathrm{DG}(b_{w^{-1}}^{d})\geq y.
Proof.

This now follows by combining the previous lemma with part (ii) of the main Lemma. ∎

In the remainder of this subsection we reprove the bounds of [Mal21, Proposition C], in the case i=1i=1:

Proposition 2.31.

Let ww be an element of a twisted Weyl group and let 𝔑\mathfrak{N} be a nimble set of roots. Then we have a sequence of inclusions

𝔑crossw(𝔑)crossw2(𝔑),\mathfrak{N}\supseteq\mathrm{cross}_{w}(\mathfrak{N})\supseteq\mathrm{cross}_{w}^{2}(\mathfrak{N})\supseteq\cdots,

which stabilises after |𝔑|(w)|\mathfrak{N}|-\ell(w) terms.

In particular we have crosswd(𝔑)=\mathrm{cross}_{w}^{d}(\mathfrak{N})=\varnothing for some integer d0d\geq 0, if and only if this holds for all d>|𝔑|(w)d>|\mathfrak{N}|-\ell(w).

Proof.

For any i0i\geq 0 let w¯i:=(w,,w,w)\underline{w}_{i}:=(w,\ldots,w,w) denote the sequence of ii copies of ww. If γ\gamma lies in crosswd(𝔑)\mathrm{cross}_{w}^{d}(\mathfrak{N}), then there exists a root β\beta in 𝔑\mathfrak{N} such that crossw¯d(γ,β)\mathrm{cross}_{\underline{w}_{d}}(\gamma,\beta) is nonempty. In other words, there exists roots γ=βd,,β0=β\gamma=\beta_{d}^{\prime},\ldots,\beta_{0}^{\prime}=\beta inductively constructed via

βj:=w(βj1+β~wβ~)\beta_{j}^{\prime}:=w(\beta_{j-1}^{\prime}+\sum_{\tilde{\beta}\in\mathfrak{R}_{w}}\tilde{\beta})

By assumption we have 𝔑crossw(𝔑)\mathfrak{N}\supseteq\mathrm{cross}_{w}(\mathfrak{N}), so β1\beta_{1}^{\prime} lies in 𝔑\mathfrak{N}. But then crossw¯d1(γ,β1)\mathrm{cross}_{\underline{w}_{d-1}}(\gamma,\beta_{1}^{\prime}) is nonempty, which means that γ\gamma also lies in crosswd1(𝔑)\mathrm{cross}_{w}^{d-1}(\mathfrak{N}). ∎

Although the sets 𝔑\mathfrak{N} and crossw(𝔑)\mathrm{cross}_{w}(\mathfrak{N}) are convex, this does not necessarily hold for the other sets appearing in such a sequence:

Example 2.32.

Consider w=s1s2s3s1w=s_{1}s_{2}s_{3}s_{1} and let v=ws2s1v=w_{\circ}s_{2}s_{1} in type 𝖡3\mathsf{B}_{3}. Then 𝔑:=v\mathfrak{N}:=\mathfrak{R}_{v} is ww-nimble and this sequence is

𝔑{α3,α23,α233}{α3,α23}{α3,α23}\mathfrak{N}\supseteq\{\alpha_{3},\alpha_{23},\alpha_{233}\}\supseteq\{\alpha_{3},\alpha_{23}\}\supseteq\{\alpha_{3},\alpha_{23}\}\supseteq\cdots
Corollary 2.33.

Let ww be an element of a twisted Weyl group. Then for any d1d\geq 1 we have

  1. (i)

    an inclusion

    wDG(bwd)+\stw,\mathfrak{R}_{w}\subseteq\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}\subseteq\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w},
  2. (ii)

    and if d>|+\stw|(w)d>|\mathfrak{R}_{+}\backslash\mathfrak{R}_{\mathrm{st}}^{w}|-\ell(w) then for any ddd^{\prime}\geq d we have

    DG(bwd)=DG(bwd).\mathrm{DG}(b_{w}^{d^{\prime}})=\mathrm{DG}(b_{w}^{d}).
Proof.

(i): If β\beta lies in stw\mathfrak{R}_{\mathrm{st}}^{w} then wd(β)w^{d}(\beta) lies in crosswd(β)\mathrm{cross}_{w}^{d}(\beta). By part (ii) of the main Lemma we have crosswd(DG(bwd))=\mathrm{cross}_{w}^{d}(\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})})=\varnothing, so β\beta does not lie in DG(bwd)\mathfrak{R}_{\mathrm{DG}(b_{w}^{d})}.

(ii): The previous proposition yields

crosswd(DG(bwd))=crosswd(DG(bwd))=\mathrm{cross}_{w}^{d}(\mathfrak{R}_{\mathrm{DG}(b_{w}^{d^{\prime}})})=\mathrm{cross}_{w}^{d^{\prime}}(\mathfrak{R}_{\mathrm{DG}(b_{w}^{d^{\prime}})})=\varnothing

and then we conclude from part (ii) of the main Lemma that DG(bwd)DG(bwd)DG(bwd)\mathrm{DG}(b_{w}^{d})\geq\mathrm{DG}(b_{w}^{d^{\prime}})\geq\mathrm{DG}(b_{w}^{d}). ∎

2.4 From roots to root subgroups

The theory of reductive groups over schemes was originally developed by Demazure and Grothendieck [SGA3-III]; some simplifications were recently made in an exposition by Conrad [Con14]. The main property that we will use is

Theorem 2.34.

Consider a split reductive group GG over a scheme. Trivialisations of the root spaces of its Lie algebra exponentiate to parametrisations pβ:𝔾aNβp_{\beta}:\mathbb{G}_{a}\overset{\sim}{\rightarrow}N_{\beta} of its root subgroups.

  1. (i)

    [Che55, p. 27] For any pair of roots β,γ\beta,\gamma with βγ\beta\neq-\gamma, there is the Chevalley commutator formula

    [pβ(cβ),pγ(cγ)]:=pβ(cβ)pγ(cγ)pβ(cβ)pγ(cγ)=i,j>0piβ+jγ(cβ,γi,jcβicγj)[p_{\beta}(c_{\beta}),p_{\gamma}(c_{\gamma})]:=p_{\beta}(c_{\beta})p_{\gamma}(c_{\gamma})p_{\beta}(-c_{\beta})p_{\gamma}(-c_{\gamma})=\prod_{i,j>0}p_{i\beta+j\gamma}(c_{\beta,\gamma}^{i,j}c_{\beta}^{i}c_{\gamma}^{j}) (2.8)

    for some global functions cβ,γi,jc_{\beta,\gamma}^{i,j} on the underlying scheme and ordering on the roots. In particular,

    NβNγ=(i0,j>0Niβ+jγ)NβN_{\beta}N_{\gamma}=\bigl{(}\prod_{i\geq 0,j>0}N_{i\beta+j\gamma}\bigr{)}N_{\beta}
  2. (ii)

    [Ste68, §6] Let w˙\dot{w} denote a lift of a Weyl group element ww to GG. Then there is the Steinberg relation

    w˙Nβ=Nw(β)w˙.\dot{w}N_{\beta}=N_{w(\beta)}\dot{w}.

In order to analyse such commutators, we will focus on the roots appearing in the product on the right-hand-side and hence define

Definition 2.35.

Let ww be an element of a twisted Weyl group. Given a convex set 𝔑\mathfrak{N} of positive roots containing w\mathfrak{R}_{w}, we set

Crossw(𝔑):=w(𝔑+w)+,\mathrm{Cross}_{w}(\mathfrak{N}):=w(\mathfrak{N}+\mathfrak{R}_{w})\cap\mathfrak{R}_{+},

and we let Crosswd(𝔑)\mathrm{Cross}_{w}^{d}(\mathfrak{N}) denote its dd-th iterate.

We won’t be using

Proposition 2.36.

Let ww be an element of a twisted Weyl group and let 𝔑\mathfrak{N} be such a set.

  1. (i)

    The sets crossw(𝔑)\mathrm{cross}_{w}(\mathfrak{N}) and Crossw(𝔑)\mathrm{Cross}_{w}(\mathfrak{N}) are also convex.

  2. (ii)

    Let 𝔏\𝔑\mathfrak{L}\subseteq\mathfrak{R}\backslash\mathfrak{N} be a subset satisfying w(𝔏)=𝔏=𝔏w(\mathfrak{L})=\mathfrak{L}=-\mathfrak{L} and such that 𝔏w\mathfrak{L}\sqcup\mathfrak{R}_{w} is convex. Then

    crossw(𝔑)𝔏=Crossw(𝔑)𝔏=.\mathrm{cross}_{w}(\mathfrak{N})\cap\mathfrak{L}=\mathrm{Cross}_{w}(\mathfrak{N})\cap\mathfrak{L}=\varnothing.
Proof.

(i): Let γ0\gamma_{0} and γ1\gamma_{1} be elements of crossw(𝔑)\mathrm{cross}_{w}(\mathfrak{N}), so γicrossw(βi)\gamma_{i}\in\mathrm{cross}_{w}(\beta_{i}^{\prime}) for some βi\beta_{i}^{\prime} in 𝔑\mathfrak{N}. If γ0+γ1\gamma_{0}+\gamma_{1} is a root, then there is a sum

w1(γ0+γ1)=w1(γ0)+w1(γ1)=β0+β1+i=1mβi,βiw,w^{-1}(\gamma_{0}+\gamma_{1})=w^{-1}(\gamma_{0})+w^{-1}(\gamma_{1})=\beta_{0}^{\prime}+\beta_{1}^{\prime}+\sum_{i=1}^{m}\beta_{i},\qquad\beta_{i}\in\mathfrak{R}_{w},

with β0\beta_{0}^{\prime} and β1\beta_{1}^{\prime} in 𝔑\mathfrak{N}. Since 𝔑\mathfrak{N} is convex and contains w\mathfrak{R}_{w}, it follows from Lemma 2.2(iii) that it contains the right-hand-side, so that crossw(𝔑)\mathrm{cross}_{w}(\mathfrak{N}) contains γ0+γ1\gamma_{0}+\gamma_{1}. The second case is analogous.

(ii): Suppose the first intersection is nonempty, so there exists a positive root β𝔑\beta\in\mathfrak{N} and a root γ𝔏\gamma\in\mathfrak{L} such that

β+i=1mβi=γ,βiw,\beta+\sum_{i=1}^{m}\beta_{i}=\gamma,\qquad\beta_{i}\in\mathfrak{R}_{w},

then as

i=1mβiγ=β\sum_{i=1}^{m}\beta_{i}-\gamma=-\beta

is a root and γ-\gamma lies in 𝔏\mathfrak{L} and w𝔏\mathfrak{R}_{w}\sqcup\mathfrak{L} is convex it again follows from Lemma 2.2(iii) that β-\beta lies in w𝔏\mathfrak{R}_{w}\sqcup\mathfrak{L}, which implies that β\beta lies in 𝔏\mathfrak{L}. The second case is analogous. ∎

These statements do not extend to higher iterates because these sets might not contain all of w\mathfrak{R}_{w}. The remainder of this subsection is devoted to proving

Lemma 2.37.

Let ww be an element of a twisted finite Weyl group and let 𝔑\mathfrak{N} be a subset of positive roots. Then

crosswd(𝔑)=if and only ifCrosswd(𝔑)=.\mathrm{cross}_{w}^{d}(\mathfrak{N})=\varnothing\qquad\textrm{if and only if}\qquad\mathrm{Cross}_{w}^{d}(\mathfrak{N})=\varnothing. (2.9)
Lemma 2.38.

Let (γ1,,γm)\sum(\gamma_{1},\ldots,\gamma_{m}) be a summing sequence of positive roots in a crystallographic root system, pick any 1km1\leq k\leq m and write γ<k:=j=1k1γj\gamma_{<k}:=\sum_{j=1}^{k-1}\gamma_{j}. Then we may partition the set {k+1,,m}\{k+1,\ldots,m\} into two subsets

{i1,,ik},{i1,,imk},\{i_{1},\ldots,i_{k^{\prime}}\},\qquad\{i_{1},\ldots,i_{m-k}\},

such that

γ<k+i=1kγijandγk+j=k+1mkγij\gamma_{<k}+\sum_{i=1}^{k^{\prime}}\gamma_{i_{j}}\qquad\textrm{and}\qquad\gamma_{k}+\sum_{j=k^{\prime}+1}^{m-k}\gamma_{i_{j}}

are both positive roots.

Proof.

We induct on mkm-k: applying Lemma 2.2(ii) to the root

γ<k+γk+γk+1,\gamma_{<k}+\gamma_{k}+\gamma_{k+1},

it follows that at least one of γ<k+γk+1\gamma_{<k}+\gamma_{k+1} or γk+γk+1\gamma_{k}+\gamma_{k+1} is a root. By replacing the corresponding pair with this sum, this also shortens the sequence and then the claim follows from the induction hypothesis. ∎

Lemma 2.39.

Let β1,,βk\beta_{1},\ldots,\beta_{k} and γ1,,γm\gamma_{1},\ldots,\gamma_{m} be positive roots in a crystallographic root system such that their sum

i=1kβi+i=1mγi=:δ\sum_{i=1}^{k}\beta_{i}+\sum_{i=1}^{m}\gamma_{i}=:\delta

is a root. Then we may partition the set {1,,m}\{1,\ldots,m\} into kk subsets {ij,1,,ij,mj}\{i_{j,1},\ldots,i_{j,m_{j}}\} with 1jk1\leq j\leq k, such that for each jj the sum

βj+l=1mjγij,l=:δj\beta_{j}+\sum_{l=1}^{m_{j}}\gamma_{i_{j,l}}=:\delta_{j}

is a root, and thus the sum over those roots yields

j=1kδj=δ.\sum_{j=1}^{k}\delta_{j}=\delta.
Proof.

We use Lemma 2.2(iii) to construct a summing sequence. Let βik\beta_{i_{k}} denote the final root from the first set of roots that appears in there, and denote the sum of the elements before it by δ<ik\delta_{<i_{k}}. Thus

(,βik,γim,,γim)=δ.\sum(\ldots,\beta_{i_{k}},\gamma_{i_{m^{\prime}}},\ldots,\gamma_{i_{m}})=\delta.

The previous lemma now yields a partition

{i1,,im′′}{im′′+1,,imm}={im,,im},\{i_{1}^{\prime},\ldots,i_{m^{\prime\prime}}^{\prime}\}\coprod\{i_{m^{\prime\prime}+1}^{\prime},\ldots,i_{m-m^{\prime}}^{\prime}\}=\{i_{m^{\prime}},\ldots,i_{m}\},

such that

δ<ik+j=1m′′γijandβik+j=m′′+1mmγij\delta_{<i_{k}}+\sum_{j=1}^{m^{\prime\prime}}\gamma_{i_{j}^{\prime}}\qquad\textrm{and}\qquad\beta_{i_{k}}+\sum_{j=m^{\prime\prime}+1}^{m-m^{\prime}}\gamma_{i_{j}^{\prime}}

are both roots and sum to δ\delta. Applying the induction hypothesis on the first sum then furnishes the claim. ∎

Lemma 2.40.

For any root βCrosswd(β)\beta^{\prime}\in\mathrm{Cross}_{w}^{d}(\beta), there exist roots β1,,βkcrosswd(β)\beta_{1}^{\prime},\ldots,\beta_{k}^{\prime}\in\mathrm{cross}_{w}^{d}(\beta) such that

i=1kβi=β.\sum_{i=1}^{k}\beta_{i}^{\prime}=\beta^{\prime}. (2.10)

In particular, (2.9) holds.

Proof.

We induct on dd, so we may assume that there exists an integer k~1\tilde{k}\in\mathbb{N}_{1} and a root β~\tilde{\beta}^{\prime} in Crosswd1(β)\mathrm{Cross}_{w}^{d-1}(\beta) such that

β=w(k~β~+i=1mγi),γiw.\beta^{\prime}=w(\tilde{k}\tilde{\beta}^{\prime}+\sum_{i=1}^{m}\gamma_{i}),\qquad\gamma_{i}\in\mathfrak{R}_{w}.

The induction hypothesis furnishes β~1,,β~kcrosswd(β)\tilde{\beta}_{1},\ldots,\tilde{\beta}_{k^{\prime}}\in\mathrm{cross}_{w}^{d}(\beta) such that i=1kβ~i=β~\sum_{i=1}^{k^{\prime}}\tilde{\beta}_{i}=\tilde{\beta}^{\prime}. Thus

w1(β)=k~i=1kβ~i+i=1mγi.w^{-1}(\beta^{\prime})=\tilde{k}\sum_{i=1}^{k^{\prime}}\tilde{\beta}_{i}+\sum_{i=1}^{m}\gamma_{i}.

The previous lemma now implies that we may rename the k~\tilde{k}-fold concatenation of β~1,,β~k\tilde{\beta}_{1},\ldots,\tilde{\beta}_{k} into β1,,βk~k\beta_{1},\ldots,\beta_{\tilde{k}k^{\prime}} and partition the γ1,,γm\gamma_{1},\ldots,\gamma_{m} such that for 1jk~k1\leq j\leq\tilde{k}k^{\prime} there are roots

β~j:=βj+γi,j,γi,jw\tilde{\beta}_{j}^{\prime}:=\beta_{j}+\sum\gamma_{i,j},\qquad\gamma_{i,j}\in\mathfrak{R}_{w}

not lying in w\mathfrak{R}_{w} (otherwise we add them to the list of γ\gamma’s and start over with a smaller list of β\beta’s). Setting βj:=w(β~j)\beta_{j}^{\prime}:=w(\tilde{\beta}_{j}^{\prime}) and k:=k~kk:=\tilde{k}k^{\prime}, we have

i=1kβj=i=1k~kw(βj+γi,j)=w(k~i=1kβ~i+i=1mγi)=β\sum_{i=1}^{k}\beta_{j}^{\prime}=\sum_{i=1}^{\tilde{k}k^{\prime}}w(\beta_{j}+\sum\gamma_{i,j})=w(\tilde{k}\sum_{i=1}^{k^{\prime}}\tilde{\beta}_{i}+\sum_{i=1}^{m}\gamma_{i})=\beta^{\prime}

so (2.10) is satisfied, and they lie in crosswd(β)\mathrm{cross}_{w}^{d}(\beta). ∎

3 Cross sections and transversality

In this section we prove part (i) and (ii) of the main Theorem.

In the first subsection we prove part (i). The main part of He-Lusztig’s proof employs the existence of certain “good” elements in each elliptic conjugacy class of the Weyl group [GM97]. Combining this part with the geometric construction of such elements in [HN12, §5.2] and unravelling the resulting proof, one finds that it is very similar to Sevostyanov’s (in the elliptic case). As He-Lusztig’s techniques are neater and yield an explicit inverse map, the proof of this subsection is based upon their approach.

More specifically, He-Lusztig constructed a candidate inverse Ψ\Psi^{\prime} to the conjugation map Ψ\Psi when the Weyl group element ww is elliptic, and proved that ΨΨ\Psi^{\prime}\circ\Psi is the identity when bwdb_{w}^{d} is divisible by bwb_{w_{\circ}} for some natural number dd [HL12, §3.7]. The core of their argument states that a certain variety with a projection map constructed out of root subgroups and sequences of Weyl group elements only depends on the image of this sequence in the braid monoid [HL12, §2.9], and this argument can be generalised to work in the nonelliptic case when LL is nontrivial. More directly however, we observe that the rôle that these Weyl group elements play here is in asserting the identity

crosswd(𝔑)=,\mathrm{cross}_{w}^{d}(\mathfrak{N})=\varnothing,

which through Lemma 2.28 is crucial in our approach to proving transversality. Rather surprisingly, the previous section demonstrated that such equations about roots are equivalent to similar identities about braids. Hence in the first subsection we’ve rewritten this part of their proof in terms of ww-crossing pairs (for arbitrary ww) satisfying this equation, yielding many new cross sections along the way.

Rather than following this up with a proof that ΨΨ\Psi\circ\Psi^{\prime} also equals the identity, He-Lusztig then appeal to Ax-Grothendieck type results about affine nn-space to conclude that Ψ\Psi^{\prime} is indeed inverse to Ψ\Psi, under suitable conditions on the base ring and its ring automorphism. However, for nonelliptic ww the slices w˙LNw\dot{w}LN_{w} are not isomorphic to affine nn-space; the following proof shows directly that ΨΨ\Psi\circ\Psi^{\prime} is the identity, shedding any conditions on the base ring and its automorphism.

In the second subsection, we will also prove that part (ii) implies the following variant on part (i):

  1. (i’)

    The conjugation action (1.4) is an isomorphism when restricted to a first order infinitesimal neighbourhood of the subscheme {id}×w˙LNw\{\mathrm{id}\}\times\dot{w}LN_{w}.

A priori (i’) is weaker; I have not studied whether they might be equivalent.

3.1 Crossing root subgroups

The following construction was inspired by [HL12, §2.7]:

Definition 3.1.

Fix an integer d1d\geq 1. We consider the set of orbits in the dd-fold Cartesian product

N(w˙L¯):=Nw˙LN××Nw˙LNN(\underline{\dot{w}L}):=N\dot{w}LN\times\cdots\times N\dot{w}LN

for the Nd1N^{d-1}-action given by

(nd1,,n1)(gd,,g1)=(gdnd1,nd11gd1nd2,,n21g2n1,n11g1).(n_{d-1},\ldots,n_{1})\cdot(g_{d},\ldots,g_{1})=(g_{d}^{\phantom{1}}n_{d-1}^{\phantom{-1}},n_{d-1}^{-1}g_{d-1}^{\phantom{1}}n_{d-2}^{\phantom{1}},\ldots,n_{2}^{-1}g_{2}^{\phantom{1}}n_{1}^{\phantom{1}},n_{1}^{-1}g_{1}^{\phantom{1}}).

We denote the (naive) orbit space by N[w˙L¯]N[\underline{\dot{w}L}] and the quotient map by

N(w˙L¯)N[w˙L¯],(gd,,g1)[gd,,g1].N(\underline{\dot{w}L})\longrightarrow N[\underline{\dot{w}L}],\qquad(g_{d},\ldots,g_{1})\longmapsto[g_{d},\ldots,g_{1}]. (3.1)

This map is equivariant with respect to the N×NN\times N-actions on N(w˙L¯)N(\underline{\dot{w}L}) and N[w˙L¯]N[\underline{\dot{w}L}] coming from left and right multiplication on the outer factors:

n(gd,,g1)n:=(ngd,,g1n),n[gd,,g1]n:=[ngd,,g1n].n^{\prime}(g_{d},\ldots,g_{1})n:=(n^{\prime}g_{d},\ldots,g_{1}n),\qquad n^{\prime}[g_{d},\ldots,g_{1}]n:=[n^{\prime}g_{d},\ldots,g_{1}n].

Given an element gNw˙LNg\in N\dot{w}LN, we shall write

[g¯]:=[g,g,,g,g]N[w˙L¯][\underline{g}]:=[g,g,\ldots,g,g]\in N[\underline{\dot{w}L}]

for the image of (g,g,,g,g)N(w˙L¯)(g,g,\ldots,g,g)\in N(\underline{\dot{w}L}) under the quotient map (3.1).

We can describe N[w˙L¯]N[\underline{\dot{w}L}] more explicitly: first consider the dd-fold product

Nw˙L¯:=w˙LNw××w˙LNwN_{\underline{\dot{w}L}}:=\dot{w}LN_{w}\times\cdots\times\dot{w}LN_{w}

and enlarge it to Nw˙L¯+:=N×Nw˙L¯N_{\underline{\dot{w}L}}^{+}:=N\times N_{\underline{\dot{w}L}}^{\,}. By multiplying the first two components of its d+1d+1-Cartesian product, we obtain a natural inclusion

Nw˙L¯+=N×w˙LNw×w˙LNw××w˙LNwNw˙LNw×w˙LNw××w˙LNw⸦⟶N(w˙L¯).N_{\underline{\dot{w}L}}^{+}=N\times\dot{w}LN_{w}\times\dot{w}LN_{w}\times\cdots\times\dot{w}LN_{w}\overset{\sim}{\longrightarrow}N\dot{w}LN_{w}\times\dot{w}LN_{w}\times\cdots\times\dot{w}LN_{w}\lhook\joinrel\longrightarrow N(\underline{\dot{w}L}). (3.2)
Notation 3.2.

We write Nw:=w1NwNN^{w}:=w^{-1}Nw\cap N for the product of root subgroups corresponding to the roots in 𝔑\w\mathfrak{N}\backslash\mathfrak{R}_{w}, and given elements x,gGx,g\in G we abbreviate left conjugation by gx:=xgx1{}^{x}g:=xgx^{-1}.

Lemma 3.3.
  1. (i)

    There is a natural factorisation

    L×NLN=LNw×L×Nw,(l,n)(n1,l,n2),L\times N\overset{\sim}{\longrightarrow}LN=L\overset{\sim}{\longrightarrow}N^{w}\times L\times N_{w},\qquad(l,n)\longmapsto(n_{1},l^{\prime},n_{2}), (3.3)

    implying Nw˙LN=Nw˙LNwN\dot{w}LN=N\dot{w}LN_{w}; if ((𝔑\w)+𝔏)𝔏=\bigl{(}(\mathfrak{N}\backslash\mathfrak{R}_{w})+\mathfrak{L}\bigl{)}\cap\mathfrak{L}=\varnothing (e.g. 𝔏\mathfrak{L} is a standard parabolic subsystem) then l=ll^{\prime}=l.

  2. (ii)

    Hence the inclusion (3.2) yields an algebraic cross section

    N(w˙L¯)\textstyle{N(\underline{\dot{w}L})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Nw˙L¯+\textstyle{N_{\underline{\dot{w}L}}^{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}N[w˙L¯]\textstyle{N[\underline{\dot{w}L}]}

    of the quotient map (3.1).

  3. (iii)

    Assume that LTL\subseteq T, pick a root β𝔑\beta\in\mathfrak{N}, elements nLNwn\in LN_{w} and mNβm\in N_{\beta}, and use (3.3) to factorise nmLNnm\in LN into a pair of elements (m1,n1)(m_{1},n_{1}) in Nw×LNwN^{w}\times LN_{w}. Then

    m1w˙γCrossw(β)NγN.{}^{\dot{w}}m_{1}\in\prod_{\gamma\in\mathrm{Cross}_{w}(\beta)}N_{\gamma}\subseteq N.
Proof.

(i): Since w𝔏\mathfrak{R}_{w}\sqcup\mathfrak{L} is convex and 𝔏=𝔏\mathfrak{L}=-\mathfrak{L} and 𝔏𝔑=\mathfrak{L}\cap\mathfrak{N}=\varnothing, it follows that

((𝔑\w)+𝔏)w=,\bigl{(}(\mathfrak{N}\backslash\mathfrak{R}_{w})+\mathfrak{L}\bigr{)}\cap\mathfrak{R}_{w}=\varnothing,

which through convexity of 𝔑𝔏\mathfrak{N}\sqcup\mathfrak{L} yields LNw=NwLLN^{w}=N^{w}L, and the first claim follows. Nimbleness then implies that

Nw˙LN=Nw˙NwLNw=Nw˙LNw.N\dot{w}LN=N\dot{w}N^{w}LN_{w}=N\dot{w}LN_{w}.

(ii) The case of d=2d=2 now follows from

Nw˙LN×𝑁Nw˙LN=Nw˙LN×𝑁Nw˙LNw=Nw˙LN×𝑁w˙LNwNw˙LN×w˙LNw=Nw˙LNw×w˙LNwN\dot{w}LN\underset{N}{\times}N\dot{w}LN=N\dot{w}LN\underset{N}{\times}N\dot{w}LN_{w}=N\dot{w}LN\underset{N}{\times}\dot{w}LN_{w}\simeq N\dot{w}LN\times\dot{w}LN_{w}=N\dot{w}LN_{w}\times\dot{w}LN_{w}

and this implies, by induction on dd, that

N[w˙L¯]=Nw˙LN×𝑁Nw˙LN×𝑁×𝑁Nw˙LN=Nw˙LNw×w˙LNw××w˙LNwNw˙L¯+.N[\underline{\dot{w}L}]=N\dot{w}LN\underset{N}{\times}N\dot{w}LN\underset{N}{\times}\cdots\underset{N}{\times}N\dot{w}LN=N\dot{w}LN_{w}\times\dot{w}LN_{w}\times\cdots\times\dot{w}LN_{w}\simeq N_{\underline{\dot{w}L}}^{+}.

(iii): Follows similarly. ∎

Notation 3.4.

We now denote by Crossw˙Ld\mathrm{Cross}_{\dot{w}L}^{d} the composition

N[w˙L¯]Nw˙L¯+=N×Nw˙L¯-↠NN[\underline{\dot{w}L}]\overset{\sim}{\longrightarrow}N_{\underline{\dot{w}L}}^{+}=N\times N_{\underline{\dot{w}L}}\relbar\joinrel\twoheadrightarrow N

of the inverse of this isomorphism with projection onto the first component of the Cartesian product.

Corollary 3.5.

Assume that LTL\subseteq T, pick an integer d>0d>0 and a positive root β\beta in 𝔑\mathfrak{N}, fix an element hN[w˙L¯]h\in N[\underline{\dot{w}L}] and consider the morphism of schemes

𝔾aNβN,mCrossw˙Ld(hm).\mathbb{G}_{a}\simeq N_{\beta}\longrightarrow N,\qquad m\longmapsto\mathrm{Cross}_{\dot{w}L}^{d}(hm).
  1. (i)

    Then

    Crossw˙Ld(hm)Crossw˙Ld(h)γCrossw(β)Nγ\mathrm{Cross}_{\dot{w}L}^{d}(hm)\in\mathrm{Cross}_{\dot{w}L}^{d}(h)\prod_{\gamma\in\mathrm{Cross}_{w}(\beta)}N_{\gamma}
  2. (ii)

    If Crossw˙Ld(h)\mathrm{Cross}_{\dot{w}L}^{d}(h) is the identity, then the derivative of this map at the identity of NβN_{\beta}

    𝔫β𝔫,xcrossw˙Ld(hx)\mathfrak{n}_{\beta}\longrightarrow\mathfrak{n},\qquad x\longmapsto\mathrm{cross}_{\dot{w}L}^{d}(hx)

    satisfies

    crossw˙Ld(hx)γcrosswd(β)𝔫γ.\mathrm{cross}_{\dot{w}L}^{d}(hx)\in\bigoplus_{\gamma\in\mathrm{cross}_{w}^{d}(\beta)}\mathfrak{n}_{\gamma}.
Proof.

(i): Denote the inverse of the element hh under the isomorphism

N×w˙LNw××w˙LNw=Nw˙L¯+N[w˙L¯]N\times\dot{w}LN_{w}\times\cdots\times\dot{w}LN_{w}=N_{\underline{\dot{w}L}}^{+}\overset{\sim}{\longrightarrow}N[\underline{\dot{w}L}]

by (m,w˙nd,,w˙n1)(m^{\prime},\dot{w}n_{d},\ldots,\dot{w}n_{1}), so mm^{\prime} lies in NN and each nin_{i} lies in LNwLN_{w}. Let (m1,n1)Nw×LNw(m_{1},n_{1}^{\prime})\in N^{w}\times LN_{w} be the factorisation of n1mLNn_{1}m\in LN in (3.3) and inductively define for 1<id1<i\leq d elements (mi,ni)Nw×LNw(m_{i},n_{i}^{\prime})\in N^{w}\times LN_{w} as the factorisation of the element ni1(mi1w˙)LNn_{i-1}({}^{\dot{w}}m_{i-1})\in LN. By induction on dd, the second part of the previous proposition implies that

miw˙γCrosswi(β)Nγ.{}^{\dot{w}}m_{i}\in\prod_{\gamma\in\mathrm{Cross}_{w}^{i}(\beta)}N_{\gamma}.

Then

hm\displaystyle hm =[m,w˙nd,,w˙n2,w˙n1m]\displaystyle=[m^{\prime},\dot{w}n_{d},\ldots,\dot{w}n_{2},\dot{w}n_{1}m]
=[m,w˙nd,,w˙n2(m1w˙),w˙n1]\displaystyle=[m^{\prime},\dot{w}n_{d},\ldots,\dot{w}n_{2}({}^{\dot{w}}m_{1}),\dot{w}n_{1}^{\prime}]
=[m(w˙md),w˙nd,,w˙n2,w˙n1]\displaystyle=[m^{\prime}(^{\dot{w}}m_{d}),\dot{w}n_{d}^{\prime},\ldots,\dot{w}n_{2}^{\prime},\dot{w}n_{1}^{\prime}]

so that

Crossw˙Ld(hm)=m(w˙md)=Crossw˙Ld(h)(w˙md).\mathrm{Cross}_{\dot{w}L}^{d}\big{(}hm\big{)}=m^{\prime}(^{\dot{w}}m_{d})=\mathrm{Cross}_{\dot{w}L}^{d}(h)(^{\dot{w}}m_{d}).

(ii): Taking derivatives with respect to the first subgroup in (2.8), the component on the right-hand-side with i>1i>1 vanishes. This implies that for d=1d=1, the image lands in

w˙(γ{β+i=1mβi+:βiw}𝔫γ)w˙1𝔫+=γcrossw(β)𝔫γ.\dot{w}\bigl{(}\bigoplus_{\gamma\in\{\beta+\sum_{i=1}^{m}\beta_{i}\in\mathfrak{R}_{+}:\beta_{i}\in\mathfrak{R}_{w}\}}\mathfrak{n}_{\gamma}\bigr{)}\dot{w}^{-1}\cap\mathfrak{n}_{+}=\bigoplus_{\gamma\in\mathrm{cross}_{w}(\beta)}\mathfrak{n}_{\gamma}.

The claim then follows by induction. ∎

Proposition 3.6.

Both crossw()\mathrm{cross}_{w}(\cdot) and Crossw()\mathrm{Cross}_{w}(\cdot) lift to the braid monoid.

Proof.

For any reduced decomposition w=uvw=uv we have

w˙Nw=u˙Nuv˙Nv\dot{w}N_{w}=\dot{u}N_{u}\dot{v}N_{v}

regardless of characteristic, and hence

βCrossw(β)Nβ×w˙Nw=w˙NwNβ=u˙Nuv˙NvNβ=βCrossu(Crossv(β))Nβ×w˙Nw,\prod_{\beta^{\prime}\in\mathrm{Cross}_{w}(\beta)}N_{\beta^{\prime}}\times\dot{w}N_{w}=\dot{w}N_{w}N_{\beta}=\dot{u}N_{u}\dot{v}N_{v}N_{\beta}=\prod_{\beta^{\prime}\in\mathrm{Cross}_{u}(\mathrm{Cross}_{v}(\beta))}N_{\beta^{\prime}}\times\dot{w}N_{w},

so that

Crossw(β)=Crossu(Crossv(β)).\mathrm{Cross}_{w}(\beta)=\mathrm{Cross}_{u}(\mathrm{Cross}_{v}(\beta)).

Taking derivatives as before then yields

crossw(β)=crossu(crossv(β)).\mathrm{cross}_{w}(\beta)=\mathrm{cross}_{u}(\mathrm{cross}_{v}(\beta)).

The rest of the proof is analogous to that of Lemma 2.6. ∎

We now prove the crucial

Lemma 3.7.

If

crosswd(β)=,\mathrm{cross}_{w}^{d}(\beta)=\varnothing,

then for all hN[w˙L¯]h\in N[\underline{\dot{w}L}] and nβNβn_{\beta}\in N_{\beta} we have

Crossw˙Ld(hnβ)=Crossw˙Ld(h).\mathrm{Cross}_{\dot{w}L}^{d}(hn_{\beta})=\mathrm{Cross}_{\dot{w}L}^{d}(h).
Proof.

Consider the notions introduced in the first paragraph of Definition 3.1; we add a tilde to denote the orbits for NLNL instead. Then from w˙L=Lw˙\dot{w}L=L\dot{w} and LNw=NwLLN_{w}=N_{w}L we similarly derive a factorisation

N×L×w˙Nw××w˙Nw=N×L×Nw¯˙=:N~w˙L¯+N~[w˙L¯]N\times L\times\dot{w}N_{w}\times\cdots\times\dot{w}N_{w}=N\times L\times N_{\underline{\dot{w}}}=:\tilde{N}_{\underline{\dot{w}L}}^{+}\overset{\sim}{\longrightarrow}\tilde{N}[\underline{\dot{w}L}]

yielding projection maps

Cross~w˙Ld:\displaystyle\widetilde{\mathrm{Cross}}_{\dot{w}L}^{d}:\quad N~[w˙L¯]N~w˙L¯+=N×L×Nw¯˙-↠N,\displaystyle\tilde{N}[\underline{\dot{w}L}]\overset{\sim}{\longrightarrow}\tilde{N}_{\underline{\dot{w}L}}^{+}=N\times L\times N_{\underline{\dot{w}}}\relbar\joinrel\twoheadrightarrow N,
projL:\displaystyle\mathrm{proj}_{L}:\quad N[w˙L¯]N~[w˙L¯]N~w˙L¯+=N×L×Nw¯˙-↠L.\displaystyle N[\underline{\dot{w}L}]\longrightarrow\tilde{N}[\underline{\dot{w}L}]\overset{\sim}{\longrightarrow}\tilde{N}_{\underline{\dot{w}L}}^{+}=N\times L\times N_{\underline{\dot{w}}}\relbar\joinrel\twoheadrightarrow L.
projw˙:\displaystyle\mathrm{proj}_{\dot{w}}:\quad N[w˙L¯]N~[w˙L¯]N~w˙L¯+=N×L×Nw¯˙-↠Nw¯˙.\displaystyle N[\underline{\dot{w}L}]\longrightarrow\tilde{N}[\underline{\dot{w}L}]\overset{\sim}{\longrightarrow}\tilde{N}_{\underline{\dot{w}L}}^{+}=N\times L\times N_{\underline{\dot{w}}}\relbar\joinrel\twoheadrightarrow N_{\underline{\dot{w}}}.
projw˙L:\displaystyle\mathrm{proj}_{\dot{w}L}:\quad N[w˙L¯]Nw˙L¯+=N×Nw˙L¯-↠Nw˙L¯.\displaystyle N[\underline{\dot{w}L}]\overset{\sim}{\longrightarrow}N_{\underline{\dot{w}L}}^{+}=N\times N_{\underline{\dot{w}L}}\relbar\joinrel\twoheadrightarrow N_{\underline{\dot{w}L}}.

This gives a natural commutative diagram

N(w˙L¯)\textstyle{N(\underline{\dot{w}L})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}N[w˙L¯]\textstyle{N[\underline{\dot{w}L}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Crossw˙Ld\scriptstyle{\mathrm{Cross}_{\dot{w}L}^{d}}N~[w˙L¯]\textstyle{\tilde{N}[\underline{\dot{w}L}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cross~w˙Ld\scriptstyle{\widetilde{\mathrm{Cross}}_{\dot{w}L}^{d}}N\textstyle{N\ignorespaces\ignorespaces\ignorespaces\ignorespaces}=\scriptstyle{=}N\textstyle{N}

Write hL:=projL(h)h_{L}:=\mathrm{proj}_{L}(h), hw˙:=projNw¯˙(h)h_{\dot{w}}:=\mathrm{proj}_{N_{\underline{\dot{w}}}}(h) and hw˙L:=projw˙L(h)h_{\dot{w}L}:=\mathrm{proj}_{\dot{w}L}(h). Then

Crossw˙Ld(hnβ)\displaystyle\mathrm{Cross}_{\dot{w}L}^{d}(hn_{\beta}) =Crossw˙Ld(h)Crossw˙Ld(hw˙Lnβ),\displaystyle=\mathrm{Cross}_{\dot{w}L}^{d}(h)\mathrm{Cross}_{\dot{w}L}^{d}(h_{\dot{w}L}n_{\beta}),
=Crossw˙Ld(h)Cross~w˙d(hLhw˙nβ)\displaystyle=\mathrm{Cross}_{\dot{w}L}^{d}(h)\widetilde{\mathrm{Cross}}_{\dot{w}}^{d}(h_{L}h_{\dot{w}}n_{\beta})
=Crossw˙Ld(h)hL1Cross~w˙d(hw˙nβ)hL\displaystyle=\mathrm{Cross}_{\dot{w}L}^{d}(h)h_{L}^{-1}\widetilde{\mathrm{Cross}}_{\dot{w}}^{d}(h_{\dot{w}}n_{\beta})h_{L}

so the claim follows from part (ii) of the previous statement. ∎

Remark 3.8.

Instead of considering orbits for NLNL, we could have also upgraded Crossw()\mathrm{Cross}_{w}(\cdot) to a version Crossw𝔏()\mathrm{Cross}_{w\mathfrak{L}}(\cdot) taking into account the roots in 𝔏\mathfrak{L}, and then proven that crosswd(β)=\mathrm{cross}_{w}^{d}(\beta)=\varnothing if and only if Crossw𝔏(β)=\mathrm{Cross}_{w\mathfrak{L}}(\beta)=\varnothing purely by studying roots. By combining this with

Crossw˙Ld(hm)Crossw˙Ld(h)γCrossw𝔏(β)Nγ,\mathrm{Cross}_{\dot{w}L}^{d}(hm)\in\mathrm{Cross}_{\dot{w}L}^{d}(h)\prod_{\gamma\in\mathrm{Cross}_{w\mathfrak{L}}(\beta)}N_{\gamma},

for LL arbitrarily, the previous lemma can be obtained almost entirely by analysing roots, but the arguments become a bit longer and perhaps less transparent, despite being essentially identical.

We can now prove (i) of the main Theorem:

Proof.

Set d>0d>0 such that (1.3) holds, then we construct the (algebraic) inverse Ψ\Psi^{\prime} to the conjugation map

Ψ:N×w˙LNwNw˙LN=Nw˙LNw,(n,w˙n~)n1w˙n~n\Psi:N\times\dot{w}LN_{w}\longrightarrow N\dot{w}LN=N\dot{w}LN_{w},\qquad(n,\dot{w}\tilde{n})\longmapsto n^{-1}\dot{w}\tilde{n}n

as follows. For an element g~Nw˙LNw\tilde{g}\in N\dot{w}LN_{w} we set

ng~:=Crossw˙Ld([g¯~])1N.n_{\tilde{g}}:=\mathrm{Cross}_{\dot{w}L}^{d}\bigl{(}[\underline{\tilde{g}}]\bigr{)}^{-1}\in N. (3.4)

Denoting the image of ng~g~ng~1Nw˙LNwn_{\tilde{g}}^{\,}\tilde{g}n_{\tilde{g}}^{-1}\in N\dot{w}LN_{w} under the inverse of the usual multiplication map

N×w˙LNwNw˙LNw,(m~,w˙n~)m~w˙n~N\times\dot{w}LN_{w}\overset{\sim}{\longrightarrow}N\dot{w}LN_{w},\qquad(\tilde{m},\dot{w}\tilde{n})\longmapsto\tilde{m}\dot{w}\tilde{n} (3.5)

by (n,gg~)(n^{\prime},g_{\tilde{g}}), now set Ψ(g~):=(ng~,gg~)\Psi^{\prime}(\tilde{g}):=(n_{\tilde{g}},g_{\tilde{g}}). We will calculate these elements more explicitly and see that equation (1.3) implies that n=idn^{\prime}=\mathrm{id}.

ΨΨ=id\Psi^{\prime}\circ\Psi=\mathrm{id}: Pick (n,g)N×w˙LNw(n,g)\in N\times\dot{w}LN_{w} and set g~:=Ψ(n,g):=n1gnNw˙LNw\tilde{g}:=\Psi(n,g):=n^{-1}gn\in N\dot{w}LN_{w}. Then

[g¯~]=[n1gn,n1gn,,n1gn]=n1[g¯]n,[\underline{\tilde{g}}]=[n^{-1}gn,n^{-1}gn,\ldots,n^{-1}gn]=n^{-1}[\underline{g}]n,

so as Crossw˙Ld([g¯])=e\mathrm{Cross}_{\dot{w}L}^{d}\bigl{(}[\underline{g}]\bigr{)}=e, Lemma 3.7 implies that

ng~1=Crossw˙Ld([g¯~])=Crossw˙Ld(n1[g¯]n)=Crossw˙Ld(n1[g¯])=Crossw˙Ld([n1g,g,,g])=n1.n_{\tilde{g}}^{-1}=\mathrm{Cross}_{\dot{w}L}^{d}\bigl{(}[\underline{\tilde{g}}]\bigl{)}=\mathrm{Cross}_{\dot{w}L}^{d}\big{(}n^{-1}[\underline{g}]n\big{)}=\mathrm{Cross}_{\dot{w}L}^{d}\big{(}n^{-1}[\underline{g}]\big{)}=\mathrm{Cross}_{\dot{w}L}^{d}\big{(}[n^{-1}g,g,\ldots,g]\big{)}=n^{-1}.

Hence ng~=nn_{\tilde{g}}=n, and thus

ng~g~ng~1=ng~n1=gn_{\tilde{g}}^{\,}\tilde{g}n_{\tilde{g}}^{-1}=n\tilde{g}n^{-1}=g

which lies in w˙LNw\dot{w}LN_{w} by assumption, so gg~=gg_{\tilde{g}}=g and therefore (ΨΨ)(n,g)=(ng~,gg~)=(n,g)(\Psi^{\prime}\circ\Psi)(n,g)=(n_{\tilde{g}},g_{\tilde{g}})=(n,g).

ΨΨ=id\Psi\circ\Psi^{\prime}=\mathrm{id}: Pick g~Nw˙LNw\tilde{g}\in N\dot{w}LN_{w} and use (3.5) to decompose g~=mw˙n\tilde{g}=m\dot{w}n for some mNm\in N and nLNwn\in LN_{w}. Let (mi,ni)Nw×LNw(m_{i},n_{i})\in N^{w}\times LN_{w} be the factorisation of nmLNnm\in LN in (3.3) for i=1i=1 and inductively for i>1i>1 as the factorisation of ni1(mi1w˙)LNn_{i-1}({}^{\dot{w}}m_{i-1})\in LN. These elements were constructed to obtain the inverse image of [g¯~][\underline{\tilde{g}}] under the isomorphism Nw˙L¯+N[w˙L¯]N_{\underline{\dot{w}L}}^{+}\overset{\sim}{\rightarrow}N[\underline{\dot{w}L}], as

[g¯~]\displaystyle[\underline{\tilde{g}}] =[mw˙n,mw˙n,,mw˙n,mw˙n]\displaystyle=[m\dot{w}n,m\dot{w}n,\ldots,m\dot{w}n,m\dot{w}n]
=[mw˙nm,w˙nm,,w˙nm,w˙n]\displaystyle=[m\dot{w}nm,\dot{w}nm,\ldots,\dot{w}nm,\dot{w}n]
=[m(w˙m1)w˙n1,(w˙m1)w˙n1,,(w˙m1)w˙n1,w˙n]\displaystyle=[m(^{\dot{w}}m_{1})\dot{w}n_{1},(^{\dot{w}}m_{1})\dot{w}n_{1},\ldots,(^{\dot{w}}m_{1})\dot{w}n_{1},\dot{w}n]
=[m((m1md1)w˙)w˙nd1,w˙nd2,,w˙n1,w˙n].\displaystyle=[m\bigl{(}{}^{\dot{w}}(m_{1}\cdots m_{d-1})\bigl{)}\dot{w}n_{d-1},\dot{w}n_{d-2},\ldots,\dot{w}n_{1},\dot{w}n].

In particular, this yields

ng~1=Crossw˙Ld([g¯~])=m((m1md1)w˙).n_{\tilde{g}}^{-1}=\mathrm{Cross}_{\dot{w}L}^{d}\big{(}[\underline{\tilde{g}}]\big{)}=m\bigl{(}{}^{\dot{w}}(m_{1}\cdots m_{d-1})\bigl{)}.

A similar calculation furnishes that md=idm_{d}=\mathrm{id}: briefly setting

a:=Crossw˙Ld([g¯~])1[g¯~]=[w˙nd1,w˙nd2,,w˙n1,w˙n]N[w˙L¯],a:=\mathrm{Cross}_{\dot{w}L}^{d}\bigl{(}[\underline{\tilde{g}}]\bigr{)}^{-1}[\underline{\tilde{g}}]=[\dot{w}n_{d-1},\dot{w}n_{d-2},\ldots,\dot{w}n_{1},\dot{w}n]\in N[\underline{\dot{w}L}],

we have by construction

am\displaystyle am =[w˙nd1,w˙nd2,,w˙n1,(w˙m1)w˙n1]\displaystyle=[\dot{w}n_{d-1},\dot{w}n_{d-2},\ldots,\dot{w}n_{1},(^{\dot{w}}m_{1})\dot{w}n_{1}]
=[w˙nd1,w˙nd2,,(w˙m2)w˙n2,w˙n1]\displaystyle=[\dot{w}n_{d-1},\dot{w}n_{d-2},\ldots,(^{\dot{w}}m_{2})\dot{w}n_{2},\dot{w}n_{1}]
=[(w˙md)w˙nd,w˙nd1,,w˙n2,w˙n1]\displaystyle=[(^{\dot{w}}m_{d})\dot{w}n_{d},\dot{w}n_{d-1},\ldots,\dot{w}n_{2},\dot{w}n_{1}]

and then Lemma 3.7 implies that

mdw˙=Crossw˙Ld(am)=Crossw˙Ld(a)=id.{}^{\dot{w}}m_{d}=\mathrm{Cross}_{\dot{w}L}^{d}(am)=\mathrm{Cross}_{\dot{w}L}^{d}(a)=\mathrm{id}.

We now obtain an expression for Ψ(g~)\Psi^{\prime}(\tilde{g}) by computing

ng~g~ng~1\displaystyle n_{\tilde{g}}\tilde{g}n_{\tilde{g}}^{-1} =((md11m11)w˙)m1mw˙nm((m1md1)w˙)\displaystyle=\bigl{(}{}^{\dot{w}}(m_{d-1}^{-1}\cdots m_{1}^{-1})\bigl{)}m^{-1}m\dot{w}nm\bigl{(}{}^{\dot{w}}(m_{1}\cdots m_{d-1})\bigl{)}
=((md11m11)w˙)w˙m1n1((m1md1)w˙)\displaystyle=\bigl{(}{}^{\dot{w}}(m_{d-1}^{-1}\cdots m_{1}^{-1})\bigl{)}\dot{w}m_{1}n_{1}\bigl{(}{}^{\dot{w}}(m_{1}\cdots m_{d-1})\bigl{)}
=((md11m11)w˙)w˙m1m2mdnd\displaystyle=\bigl{(}{}^{\dot{w}}(m_{d-1}^{-1}\cdots m_{1}^{-1})\bigl{)}\dot{w}m_{1}m_{2}\cdots m_{d}n_{d}
=w˙mdnd=w˙nd\displaystyle=\dot{w}m_{d}n_{d}=\dot{w}n_{d}

which already lies in w˙LNw\dot{w}LN_{w}. Thus gg~=w˙nd=ng~g~ng~1g_{\tilde{g}}=\dot{w}n_{d}=n_{\tilde{g}}^{\,}\tilde{g}n_{\tilde{g}}^{-1}, and hence (ΨΨ)(g~)=Ψ(ng~,ng~g~ng~1)=g~(\Psi\circ\Psi^{\prime})(\tilde{g})=\Psi(n_{\tilde{g}}^{\,},n_{\tilde{g}}^{\,}\tilde{g}n_{\tilde{g}}^{-1})=\tilde{g}. ∎

Remark 3.9.
  1. (i)

    We could have written the same proof with N~(w˙L¯)\tilde{N}(\underline{\dot{w}L}) (as defined in the proof of Lemma 3.7) instead of N(w˙L¯)N(\underline{\dot{w}L}); nothing changes except for the final paragraph, where more factorising is required.

  2. (ii)

    For convenience, let’s briefly add a dd to the notions introduced in Definition 3.1, so Nd(w˙L¯):=N(w˙L¯){}_{d}N(\underline{\dot{w}L}):=N(\underline{\dot{w}L}), etc. The previous proof easily generalises to show that if (1.3) holds for some dd, then

    N×Nw˙L¯iNi[w˙L¯],(n,[gi,,g1])[n1gi,,g1n]N\times{}_{i}N_{\underline{\dot{w}L}}^{\,}\longrightarrow{}_{i}N[\underline{\dot{w}L}],\qquad\bigl{(}n,[g_{i},\ldots,g_{1}]\bigr{)}\longmapsto[n^{-1}g_{i},\ldots,g_{1}n]

    is an isomorphism for any i1i\geq 1; this proof was just the case i=1i=1.

  3. (iii)

    If 𝔏\mathfrak{L} is a standard parabolic subsystem then Lemma 3.3(i) implies that the LL-component of image of nw˙lnNw˙LNn\dot{w}ln^{\prime}\in N\dot{w}LN in w˙LN\dot{w}LN under the inverse map Ψ\Psi^{\prime} is again ll; this plays a rôle in [Sev19, Lemma 5.7].

3.2 Charts on the quotient stack [G/G][G/G]

Sevostyanov deduced from the cross section isomorphism (1.4) that his slices transversely intersect the conjugacy classes of GG [Sev11, Proposition 2.3]. In this subsection we adapt his approach to prove that (ii) still holds in our more general setting, and simultaneously refine it to show that (ii) \Rightarrow (i’):

Notation 3.10.

We denote by Adm()\mathrm{Ad}_{m}(\cdot) the right adjoint action map of an element mm of GG on its Lie algebra 𝔤\mathfrak{g}. We let 𝔫w,𝔫w,𝔩\mathfrak{n}_{w},\mathfrak{n}^{w},\mathfrak{l}, 𝔫¯\mathfrak{\overline{n}}, 𝔫¯w\overline{\mathfrak{n}}^{w}, 𝔫¯w1\mathfrak{\overline{n}}_{w^{-1}} denote the free submodules of 𝔤\mathfrak{g} corresponding to roots in w,𝔑\w,𝔏\mathfrak{R}_{w},\mathfrak{N}\backslash\mathfrak{R}_{w},\mathfrak{L}, 𝔑-\mathfrak{N}, (𝔑\w)-(\mathfrak{N}\backslash\mathfrak{R}_{w}) and w(w)=w1w(\mathfrak{R}_{w})=-\mathfrak{R}_{w^{-1}} respectively. (These are actually all Lie subalgebras, as convexity of +\w\mathfrak{R}_{+}\backslash\mathfrak{R}_{w} and 𝔑\mathfrak{N} implies that 𝔑\w\mathfrak{N}\backslash\mathfrak{R}_{w} is convex.) We let 𝔱w\mathfrak{t}_{w}^{\prime} denote the orthogonal complement inside 𝔱\mathfrak{t} to 𝔩𝔱\mathfrak{l}\cap\mathfrak{t}; since LL contains TwT^{w} we have 𝔱w𝔱w=(𝔱w)\mathfrak{t}_{w}^{\prime}\subseteq\mathfrak{t}_{w}=(\mathfrak{t}^{w})^{\bot}, which is ww-invariant as both TwT^{w} and 𝔏\mathfrak{L} are. Finally, we denote by N¯\overline{N} the unipotent subgroup of GG corresponding to 𝔫¯\overline{\mathfrak{n}} (and 𝔑-\mathfrak{N}).

Lemma 3.11.

The image of the differential of the conjugation map

G×w˙LNwG,(g,m)g1mgG\times\dot{w}LN_{w}\longrightarrow G,\qquad(g,m)\longmapsto g^{-1}mg (3.6)

at any point (id,m)(\mathrm{id},m) is given in the left trivalisation of the tangent bundle of GG by

(idAdm)(𝔫𝔫¯)+𝔱w𝔩𝔫w𝔫¯w1.(\mathrm{id}-\mathrm{Ad}_{m})(\mathfrak{n}\oplus\overline{\mathfrak{n}})+\mathfrak{t}_{w}^{\prime}\oplus\mathfrak{l}\oplus\mathfrak{n}_{w}\oplus\overline{\mathfrak{n}}_{w^{-1}}.

Moreover, we have

(idAdm)(𝔫¯)𝔫¯𝔩𝔫w.(\mathrm{id}-\mathrm{Ad}_{m})(\overline{\mathfrak{n}})\subseteq\overline{\mathfrak{n}}\oplus\mathfrak{l}\oplus\mathfrak{n}_{w}.
Proof.

The left trivialisation of the tangent bundle of GG induces for all points mm in w˙LNw\dot{w}LN_{w} identifications of their tangent spaces Tm(w˙LNw)𝔩𝔫wT_{m}(\dot{w}LN_{w})\simeq\mathfrak{l}\oplus\mathfrak{n}_{w} with a free submodule of 𝔤\mathfrak{g}, and this differential is then given at a point (id,m)(\mathrm{id},m) by the linear map

𝔤(𝔩𝔫w)𝔤,(x,l)(idAdm)(x)+l.\mathfrak{g}\oplus(\mathfrak{l}\oplus\mathfrak{n}_{w})\longrightarrow\mathfrak{g},\qquad(x,l)\longmapsto(\mathrm{id}-\mathrm{Ad}_{m})(x)+l. (3.7)

For any t𝔱t\in\mathfrak{t} and mm^{\prime} in LL we have Adm(t)t+𝔩\mathrm{Ad}_{m^{\prime}}(t)\in t+\mathfrak{l}. As w𝔏\mathfrak{R}_{w}\cup\mathfrak{L} is convex and 𝔏=𝔏\mathfrak{L}=-\mathfrak{L}, it then follows for m′′Nwm^{\prime\prime}\in N_{w} that Adm′′(t+𝔩)t+𝔩𝔫w\mathrm{Ad}_{m^{\prime\prime}}(t+\mathfrak{l})\in t+\mathfrak{l}\oplus\mathfrak{n}_{w}, so that finally for m=w˙mm′′m=\dot{w}m^{\prime}m^{\prime\prime} we have

Adm(t)=Adm′′AdmAdw(t)Adw(t)+𝔩𝔫w.\mathrm{Ad}_{m}(t)=\mathrm{Ad}_{m^{\prime\prime}}\mathrm{Ad}_{m^{\prime}}\mathrm{Ad}_{w}(t)\in\mathrm{Ad}_{w}(t)+\mathfrak{l}\oplus\mathfrak{n}_{w}.

By definition of 𝔱w\mathfrak{t}_{w} the linear operator idw\mathrm{id}-w restricts to an isomorphism, and as 𝔱w\mathfrak{t}_{w}^{\prime} is invariant under ww it then further restricts to an isomorphism of 𝔱w\mathfrak{t}_{w}^{\prime}. Hence the previous equation now yields

(idAdm)(𝔱w)+𝔩𝔫w=(idAdw)(𝔱w)+𝔩𝔫w=𝔱w𝔩𝔫w.(\mathrm{id}-\mathrm{Ad}_{m})(\mathfrak{t}_{w}^{\prime})+\mathfrak{l}\oplus\mathfrak{n}_{w}=(\mathrm{id}-\mathrm{Ad}_{w})(\mathfrak{t}_{w}^{\prime})+\mathfrak{l}\oplus\mathfrak{n}_{w}=\mathfrak{t}_{w}^{\prime}\oplus\mathfrak{l}\oplus\mathfrak{n}_{w}.

Similarly, from Adm(𝔩)=Adm′′(𝔩)𝔩𝔫w\mathrm{Ad}_{m}(\mathfrak{l})=\mathrm{Ad}_{m^{\prime\prime}}(\mathfrak{l})\subseteq\mathfrak{l}\oplus\mathfrak{n}_{w} it follows that

(idAdm)(𝔩)+𝔩𝔫w=𝔩𝔫w.(\mathrm{id}-\mathrm{Ad}_{m})(\mathfrak{l})+\mathfrak{l}\oplus\mathfrak{n}_{w}=\mathfrak{l}\oplus\mathfrak{n}_{w}.

Finally, from Adm(𝔫¯w1)=Adm′′Adm(𝔫w)𝔩𝔫w\mathrm{Ad}_{m}(\overline{\mathfrak{n}}_{w^{-1}})=\mathrm{Ad}_{m^{\prime\prime}}\mathrm{Ad}_{m^{\prime}}(\mathfrak{n}_{w})\subseteq\mathfrak{l}\oplus\mathfrak{n}_{w} we deduce that

(idAdm)(𝔫¯w1)+𝔩𝔫w=𝔩𝔫w𝔫¯w1.(\mathrm{id}-\mathrm{Ad}_{m})(\overline{\mathfrak{n}}_{w^{-1}})+\mathfrak{l}\oplus\mathfrak{n}_{w}=\mathfrak{l}\oplus\mathfrak{n}_{w}\oplus\overline{\mathfrak{n}}_{w^{-1}}.

Since the pair (𝔑,𝔏)(\mathfrak{N},\mathfrak{L}) is slicing there is a decomposition 𝔤𝔫𝔫¯𝔱w𝔩\mathfrak{g}\simeq\mathfrak{n}\oplus\overline{\mathfrak{n}}\oplus\mathfrak{t}_{w}^{\prime}\oplus\mathfrak{l}, so the first claim now follows.

As 𝔏w\mathfrak{L}\cup\mathfrak{R}_{w} and 𝔏𝔑\mathfrak{L}\cup\mathfrak{N} are convex, it follows from 𝔏=𝔏\mathfrak{L}=-\mathfrak{L} that so is 𝔏(𝔑\w)\mathfrak{L}\cup(\mathfrak{N}\backslash\mathfrak{R}_{w}). But then 𝔏(𝔑\w)\mathfrak{L}\cup-(\mathfrak{N}\backslash\mathfrak{R}_{w}) is also convex, so that Adm(𝔫¯w)𝔫¯w𝔩\mathrm{Ad}_{m^{\prime}}(\overline{\mathfrak{n}}^{w})\subseteq\overline{\mathfrak{n}}^{w}\oplus\mathfrak{l} for any mm^{\prime} in LL. As

(𝔑\w)w=(𝔑\w)w=,-(\mathfrak{N}\backslash\mathfrak{R}_{w})\cap-\mathfrak{R}_{w}=(\mathfrak{N}\backslash\mathfrak{R}_{w})\cap\mathfrak{R}_{w}=\varnothing,

for any m′′m^{\prime\prime} in NwN_{w} the operator Adm′′\mathrm{Ad}_{m^{\prime\prime}} acts as the identity on 𝔫¯w\overline{\mathfrak{n}}^{w}. Then for m=w˙mm′′m=\dot{w}m^{\prime}m^{\prime\prime} we have

Adm(𝔫¯)=Adm′′AdmAdw(𝔫¯)=Adm′′Adm(𝔫¯w𝔫w)Adm′′(𝔫¯w𝔩𝔫w)=𝔫¯w𝔩𝔫w,\mathrm{Ad}_{m}(\overline{\mathfrak{n}})=\mathrm{Ad}_{m^{\prime\prime}}\mathrm{Ad}_{m^{\prime}}\mathrm{Ad}_{w}(\overline{\mathfrak{n}})=\mathrm{Ad}_{m^{\prime\prime}}\mathrm{Ad}_{m^{\prime}}(\overline{\mathfrak{n}}^{w}\oplus\mathfrak{n}_{w})\subseteq\mathrm{Ad}_{m^{\prime\prime}}(\overline{\mathfrak{n}}^{w}\oplus\mathfrak{l}\oplus\mathfrak{n}_{w})=\overline{\mathfrak{n}}^{w}\oplus\mathfrak{l}\oplus\mathfrak{n}_{w},

yielding the second claim. ∎

We now prove (ii):

Proof.

Since GG and w˙LNw\dot{w}LN_{w} are smooth, the claim is equivalent to requiring that the image of the differential of (3.6) is surjective at each point of G×w˙LNwG\times\dot{w}LN_{w}. By equivariance for the GG-action on the first component by left translation, it suffices to prove this at each point of the form (id,m)(\mathrm{id},m). When we restrict (3.6) to NN, it yields the cross section morphism (1.4) which is an isomorphism by the previous subsection. In the left trivialisation we have

𝔫Adm(𝔫w1)𝔩𝔫Tm(Nw1w˙LN)=Tm(Nw˙LN),\mathfrak{n}\subseteq\mathrm{Ad}_{m}(\mathfrak{n}_{w^{-1}})\oplus\mathfrak{l}\oplus\mathfrak{n}\simeq T_{m}(N_{w^{-1}}\dot{w}LN)=T_{m}(N\dot{w}LN),

so by this isomorphism the image of the differential certainly contains 𝔫\mathfrak{n}.

Now consider the Chevalley anti-automorphism which switches positive and negative root vectors of a Chevalley basis. Expressing a lift of a simple reflection as a product of exponentials of such elements, an SL2\mathrm{SL}_{2}-calculation shows that its image under this involution is again a lift of this simple reflection. Hence the involution maps a lift of w1w^{-1} to a lift of ww, but then the image of the slice for (a suitable lift of) w1w^{-1} is

N¯w1L¯w˙=N¯w1Lw˙=N¯w1w˙L=w˙NwL=w˙LNw.\overline{N}_{w^{-1}}\overline{L}\dot{w}=\overline{N}_{w^{-1}}L\dot{w}=\overline{N}_{w^{-1}}\dot{w}L=\dot{w}N_{w}L=\dot{w}LN_{w}.

Since by assumption equation (1.3) holds and the pair is slicing, Lemma 2.28 implies that equation (1.3) also holds with ww replaced by w1w^{-1}. Thus by the previous subsection the cross section isomorphism (1.4) also holds for w1w^{-1}. Hence from the involution we now obtain an isomorphism

N¯×w˙LNwN¯w˙LN¯,\overline{N}\times\dot{w}LN_{w}\overset{\sim}{\longrightarrow}\overline{N}\dot{w}L\overline{N},

so by the same reasoning as in the previous paragraph, the image of the differential (in the left trivialisation) also contains 𝔫¯\overline{\mathfrak{n}}. Combining this with the first part of the previous lemma, the claim again follows from the decomposition 𝔤𝔫𝔫¯𝔱w𝔩\mathfrak{g}\simeq\mathfrak{n}\oplus\overline{\mathfrak{n}}\oplus\mathfrak{t}_{w}^{\prime}\oplus\mathfrak{l}. ∎

And finally, we prove (ii) \Rightarrow (i’):

Proof.

Concretely, (i’) says that image of the differential of the conjugation map

N×w˙LNwNw˙LN,(g,m)g1mgN\times\dot{w}LN_{w}\longrightarrow N\dot{w}LN,\qquad(g,m)\longmapsto g^{-1}mg

at any point (id,m)(\mathrm{id},m) is an isomorphism. In the left trivialisation we have

Tm(Nw˙LN)=Tm(Nw˙LNw)Adm(𝔫)𝔩𝔫wT_{m}(N\dot{w}LN)=T_{m}(N\dot{w}LN_{w})\simeq\mathrm{Ad}_{m}(\mathfrak{n})\oplus\mathfrak{l}\oplus\mathfrak{n}_{w}

and this differential is given by

𝔫(𝔩𝔫w)Adm(𝔫)𝔩𝔫w,(x,l)(idAdm)(x)+l.\mathfrak{n}\oplus(\mathfrak{l}\oplus\mathfrak{n}_{w})\longrightarrow\mathrm{Ad}_{m}(\mathfrak{n})\oplus\mathfrak{l}\oplus\mathfrak{n}_{w},\qquad(x,l)\longmapsto(\mathrm{id}-\mathrm{Ad}_{m})(x)+l. (3.8)

By assumption the differential of (3.6) is surjective. Since (𝔫¯𝔫w)𝔫w=(\overline{\mathfrak{n}}\oplus\mathfrak{n}_{w})\cap\mathfrak{n}^{w}=\varnothing, the decomposition 𝔤=𝔫w𝔫¯𝔩𝔫w𝔱w\mathfrak{g}=\mathfrak{n}^{w}\oplus\overline{\mathfrak{n}}\oplus\mathfrak{l}\oplus\mathfrak{n}_{w}\oplus\mathfrak{t}_{w}^{\prime} and the statements of the previous lemma imply that 𝔫w(idAdm)(𝔫)+𝔩𝔫w\mathfrak{n}^{w}\subseteq(\mathrm{id}-\mathrm{Ad}_{m})(\mathfrak{n})+\mathfrak{l}\oplus\mathfrak{n}_{w}. But then we have

𝔫(idAdm)(𝔫)+𝔩𝔫wAdm(𝔫)𝔩𝔫w\mathfrak{n}\subseteq(\mathrm{id}-\mathrm{Ad}_{m})(\mathfrak{n})+\mathfrak{l}\oplus\mathfrak{n}_{w}\subseteq\mathrm{Ad}_{m}(\mathfrak{n})\oplus\mathfrak{l}\oplus\mathfrak{n}_{w}

which implies that the second inclusion is an equality. Hence the differential (3.8) is surjective; as it is a morphism of (sheaves of) (locally) free modules of finite rank, it is thus an isomorphism. ∎

4 Poisson reduction

Having obtained that the action of NN on Nw˙LNN\dot{w}LN is free (whilst working over \mathbb{C}, for his particular choice of firmly convex Weyl group element ww with the slicing pair (+\w,w)(\mathfrak{R}_{+}\backslash\mathfrak{R}^{w},\mathfrak{R}^{w})), Sevostyanov proceeds to proving that the Semenov-Tian-Shansky bracket on GG reduces to the slice w˙LNw\dot{w}LN_{w} when the rr-matrix is changed from the standard one rstr_{\mathrm{st}} to

r=rst+1+w1wproj𝔱w=proj𝔫++1+w1wproj𝔱wproj𝔫,r=r_{\mathrm{st}}+\frac{1+w}{1-w}\mathrm{proj}_{\mathfrak{t}_{w}}=\mathrm{proj}_{\mathfrak{n}_{+}}+\frac{1+w}{1-w}\mathrm{proj}_{\mathfrak{t}_{w}}-\mathrm{proj}_{\mathfrak{n}_{-}},

by employing a general Poisson reduction method for manifolds [MR86, §2]. This approach was based on earlier work he did with his advisor on their loop analogues [STSS98, Theorem 2.5]. We will continue to work in the algebraic setting:

Definition 4.1.

A Poisson scheme is a scheme XX with a Poisson bracket on its sheaf of functions. On its smooth locus this bracket corresponds to a Poisson bivector field which we will denote by Π\Pi; there it induces a musical morphism Π#:TXTX\Pi^{\#}:T^{*}X\rightarrow TX from the cotangent sheaf to the tangent sheaf. A function ff then defines a Hamiltonian vector field Hamf=Π#(df)\mathrm{Ham}_{f}=\Pi^{\#}(\mathrm{d}f) on this locus.

We modify this reduction method in Proposition 4.25; one obtains a statement that is very similar to a standard characterisation of smooth Poisson subschemes (recalled in Proposition 4.6), which explains the focus on Hamiltonian vector fields in the final proof. By analysing tangent spaces with some new root combinatorics we can work with a larger class of factorisable rr-matrices, and settle which of them yield reducible Poisson brackets.

As in the previous sections, we are implicitly working over a base scheme but will omit it from all notation.

4.1 Coisotropic subgroups

Motivated by work of physicists on integrable systems, Drinfeld initiated the study of Poisson-Lie groups (and their quantisations); they translate to the algebraic setting as

Definition 4.2 ([Dri83, §3]).

A group scheme GG equipped with a Poisson bracket is called a Poisson algebraic group if this bracket is multiplicative, i.e. if the multiplication map G×GGG\times G\rightarrow G is a morphism of Poisson schemes.

The identity element of GG is then a symplectic point, so that the Poisson bracket

𝒪G,id𝒪G,id𝒪G,id,ff(dfdf)(Π)\mathcal{O}_{G,\mathrm{id}}\otimes\mathcal{O}_{G,\mathrm{id}}\longrightarrow\mathcal{O}_{G,\mathrm{id}},\qquad f\otimes f^{\prime}\longmapsto(\mathrm{d}f\otimes\mathrm{d}f^{\prime})(\Pi)

induces on its tangent space the structure of its Lie bialgebra.

Semenov-Tian-Shansky used the formalism of Poisson algebraic groups to study the “hidden symmetry groups” (dressing transformations) of certain integrable systems, as these don’t preserve Poisson structures; instead, they are Poisson actions:

Definition 4.3 ([STS85, p. 1238]).

A group action of a Poisson algebraic group GG on a Poisson scheme XX is called Poisson if the action map G×XXG\times X\rightarrow X is a morphism of Poisson schemes.

Concretely, a point xx in XX then induces a map x:GXx:G\rightarrow X via ggxg\mapsto gx and in terms of the Poisson brackets on GG and XX, the Poisson condition can then be rephrased as

{f,f}(gx)={xf,xf}(g)+{gf,gf}(x)\{f,f^{\prime}\}(gx)=\{x^{*}f,x^{*}f^{\prime}\}(g)+\{g^{*}f,g^{*}f^{\prime}\}(x) (4.1)

for f,f𝒪Xf,f^{\prime}\in\mathcal{O}_{X} and arbitrary points gGg\in G, xXx\in X.

As shown at the end of this subsection, in order to construct interesting quotients out of Poisson actions one sometimes uses subgroups of GG that are not necessarily Poisson themselves:

Definition 4.4.

Let XX be a Poisson scheme. A smooth closed subscheme ZXZ\hookrightarrow X is called coisotropic (resp. Poisson) if

ΠzTzZTzX(resp. ΠzTzZTzZ)\Pi_{z}\in T_{z}Z\wedge T_{z}X\qquad(\textrm{resp. }\Pi_{z}\in T_{z}Z\wedge T_{z}Z)

for all points zz lying in ZZ.

Notation 4.5.

Given a scheme XX we denote its sheaf of functions by 𝒪X\mathcal{O}_{X}. The inclusion of a closed subscheme ι:ZX\iota:Z\hookrightarrow X induces a morphism ι1(𝒪X)𝒪Z\iota^{-1}(\mathcal{O}_{X})\rightarrow\mathcal{O}_{Z}, and its ideal ideal sheaf is denoted by Z:=ker(ι:ι1(𝒪X)𝒪Z)\mathcal{I}_{Z}:=\mathrm{ker}\bigl{(}\iota^{\sharp}:\iota^{-1}(\mathcal{O}_{X})\rightarrow\mathcal{O}_{Z}\bigr{)}. Given a function ff in ι1(𝒪X)\iota^{-1}(\mathcal{O}_{X}) we denote its image under ι\iota^{\sharp} by f|Zf|_{Z}.

The focus in the final proof of this section will be on Hamiltonian vector fields; heuristically, this is due to a group action analogue of the following

Proposition 4.6 ([Wei83, Lemma 1.1]).

Let XX be a Poisson scheme and ι:ZX\iota:Z\hookrightarrow X a smooth closed subscheme. Then the following are equivalent:

  1. (i)

    ZZ is Poisson.

  2. (ii)

    ker(ι1(𝒪X)𝒪Z)\mathrm{ker}\bigl{(}\iota^{-1}(\mathcal{O}_{X})\twoheadrightarrow\mathcal{O}_{Z}\bigr{)} is a subsheaf of Poisson ideals; in other words, the Poisson bracket on 𝒪X\mathcal{O}_{X} reduces to 𝒪Z\mathcal{O}_{Z}.

  3. (iii)

    For any function ff in ι1(𝒪X)\iota^{-1}(\mathcal{O}_{X}), its Hamiltonian vector field lies in TZTX|ZT_{Z}\subseteq T_{X}|_{Z}.

An algebraic proof of (ii) \Leftrightarrow (iii) can be recovered as a special case of Lemma 4.25. One can characterise coisotropic smooth closed subschemes similarly [Wei88, Proposition 1.2.2].

Proposition 4.7.

Let GG be a Poisson algebraic group and let HH be a closed algebraic subgroup.

  1. (i)

    If HH is coisotropic, then the annihilator 𝔥0𝔤\mathfrak{h}^{0}\subseteq\mathfrak{g}^{*} of its Lie algebra 𝔥𝔤\mathfrak{h}\subseteq\mathfrak{g} is a Lie subalgebra of 𝔤\mathfrak{g}^{*}.

  2. (ii)

    [STS85, Proposition 2] If HH is Poisson, then this annihilator 𝔥0\mathfrak{h}^{0} is an ideal of 𝔤\mathfrak{g}^{*}.

If HH is connected, then the converse to (i) and (ii) holds as well.

Proof.

(i) is well-known, it follows from: let VV be a locally free module over a ring, Π\Pi an element of VVV\wedge V and UU a locally free submodule of VV; denote the annihilator of UU in the dual VV^{*} by U0U^{0}. Then Π#(U0)U\Pi^{\#}(U^{0})\subseteq U if and only if ΠUV\Pi\in U\wedge V.

(ii): Similarly, this follows from Π#(V)U\Pi^{\#}(V^{*})\subseteq U if and only if ΠUU\Pi\in U\wedge U. ∎

Notation 4.8.

If a group scheme HH acts on a scheme XX, then we denote the resulting sheaf of HH-invariant functions on XX by 𝒪XH\mathcal{O}_{X}^{H}.

Proposition 4.9 ([STS85, Theorem 6]).

Let GG be a Poisson algebraic group with a Poisson action on a Poisson scheme XX, and let HH be a closed coisotropic subgroup of GG. Furthermore assume that the restriction of the action on XX to HH preserves a closed subscheme ι:ZX\iota:Z\hookrightarrow X. Then (ι)1(𝒪ZH)(\iota^{\sharp})^{-1}(\mathcal{O}_{Z}^{H}) is a sheaf of Poisson subalgebras of ι1(𝒪X)\iota^{-1}(\mathcal{O}_{X}).

Proof.

Let f,f(ι)1(𝒪ZH)f,f^{\prime}\in(\iota^{\sharp})^{-1}(\mathcal{O}_{Z}^{H}), hHh\in H and let zZz\in Z. Since HH is coisotropic we have

{zf,zg}(h)=(zΠh)(f,g)=0,\{z^{*}f,z^{*}g\}(h)=(z_{*}\Pi_{h})(f,g)=0,

so as the action is Poisson it then follows from (4.1) that

{f,f}(hz)={zf,zf}(h)+{hf,hf}(z)={hf,hf}(z)={f,f}(z).\{f,f^{\prime}\}(hz)=\{z^{*}f,z^{*}f^{\prime}\}(h)+\{h^{*}f,h^{*}f^{\prime}\}(z)=\{h^{*}f,h^{*}f^{\prime}\}(z)=\{f,f^{\prime}\}(z).\qed

4.2 Factorisable rr-matrices

In order to obtain explicit coisotropic subgroups for reductive group schemes, we will now specialise to a particular class of Poisson structures.

Notation 4.10.

Given an element x=i=1naibix=\sum_{i=1}^{n}a_{i}\otimes b_{i} in 𝔤𝔤\mathfrak{g}\otimes\mathfrak{g}, we will write x12:=i=1naibi1x_{12}:=\sum_{i=1}^{n}a_{i}\otimes b_{i}\otimes 1 and x13:=i=1nai1bix_{13}:=\sum_{i=1}^{n}a_{i}\otimes 1\otimes b_{i} and similarly x23x_{23} for the usual elements in 3𝔤\otimes^{3}\mathfrak{g}. Furthermore, we denote its “flip” by x21:=i=1nbiai𝔤𝔤x^{21}:=\sum_{i=1}^{n}b_{i}\otimes a_{i}\in\mathfrak{g}\otimes\mathfrak{g}.

The existence of a multiplicative Poisson structure can be rephrased cohomologically: using the left or right trivialisation, they are 1-cocycles on GG with values in 𝔤𝔤\mathfrak{g}\wedge\mathfrak{g}. Subsequently employing e.g. Whitehead’s first lemma, there often exists an element r𝔤𝔤r\in\mathfrak{g}\wedge\mathfrak{g} such that the Lie cobracket δ\delta of a Lie bialgebra 𝔤\mathfrak{g} equals the differential r\partial r, i.e. such that

δ(x)=r(x):=12[x1+1x,r]\delta(x)=\partial r(x):=\frac{1}{2}[x\otimes 1+1\otimes x,r]

for all x𝔤x\in\mathfrak{g}. Conversely, in order for an arbitrary element r𝔤𝔤r\in\mathfrak{g}\wedge\mathfrak{g} to define a compatible cobracket it is necessary and sufficient [Dri83, §6] that this rr-matrix satisfies the generalised Yang-Baxter equation

[r,r](3𝔤)𝔤,[r,r]\in(\wedge^{3}\mathfrak{g})^{\mathfrak{g}},

where for any element r𝔤𝔤r\in\mathfrak{g}\otimes\mathfrak{g} we denote by

[r,r]:=[r12,r13]+[r12,r23]+[r13,r23]3𝔤[r,r]:=[r_{12},r_{13}]+[r_{12},r_{23}]+[r_{13},r_{23}]\in\otimes^{3}\mathfrak{g}

its Drinfeld bracket; up to scalar, this coincides with the canonical Gerstenhaber (or Schouten-Nijenhuis) bracket when restricted to 2𝔤𝔤\wedge^{2}\mathfrak{g}\subset\wedge^{\bullet}\mathfrak{g}. Note that for any element cc in (𝔤𝔤)𝔤(\mathfrak{g}\otimes\mathfrak{g})^{\mathfrak{g}}, we have [c,c]=[c12,c23][c,c]=[c_{12},c_{23}].

Theorem 4.11.

Let 𝔤\mathfrak{g} be a reductive Lie algebra over a field 𝕜\Bbbk.

  1. (i)

    [CE48, Theorem 19.1] There exists a natural quasi-isomorphism between the exterior algebra (𝔤)𝔤(\wedge^{\bullet}\mathfrak{g})^{\mathfrak{g}} and the Chevalley-Eilenberg complex computing Lie algebra homology, and hence

    (𝔤)𝔤H(𝔤).(\wedge^{\bullet}\mathfrak{g})^{\mathfrak{g}}\simeq\mathrm{H}_{\bullet}(\mathfrak{g}).
  2. (ii)

    [Kos50, §11] If the characteristic is 0 (or sufficiently large) then the Koszul map

    Sym2(𝔤)𝔤(3𝔤)𝔤,c[c12,c13]\mathrm{Sym}^{2}(\mathfrak{g})^{\mathfrak{g}}\longrightarrow(\wedge^{3}\mathfrak{g})^{\mathfrak{g}},\qquad c\longmapsto[c_{12},c_{13}]

    is an isomorphism.

Throughout the rest of this section, we will implicitly make use of the identification

𝔤𝔤𝔤𝔤Hom(𝔤,𝔤),xy(ξξ(x)y).\mathfrak{g}\wedge\mathfrak{g}\subset\mathfrak{g}\otimes\mathfrak{g}\simeq\mathrm{Hom}(\mathfrak{g}^{*},\mathfrak{g}),\qquad x\otimes y\longmapsto\bigl{(}\xi\mapsto\xi(x)y\bigr{)}.
Proposition 4.12 ([BD82]).

Let 𝔤\mathfrak{g} be a Lie algebra over a ring with 2 invertible, let r𝔤𝔤r\in\mathfrak{g}\wedge\mathfrak{g} and let cSym2(𝔤)𝔤c\in\mathrm{Sym}^{2}(\mathfrak{g})^{\mathfrak{g}} and consider the two maps

r±:𝔤𝔤,ξr±c2ξ.r_{\pm}:\mathfrak{g}^{*}\longrightarrow\mathfrak{g},\qquad\xi\longmapsto\frac{r\pm c}{2}\xi.

Then the following are equivalent:

  1. (i)

    [r,r]=[c,c][r,r]=-[c,c].

  2. (ii)

    The map r±r_{\pm} is a Lie algebra homomorphism.

  3. (iii)

    The map r±r_{\pm} is a Lie coalgebra antihomomorphism.

The equation [r,r]=[c,c][r,r]=-[c,c] is called the modified classical Yang-Baxter equation.

Definition 4.13.

If GG is a Poisson algebraic group whose Lie bialgebra 𝔤\mathfrak{g} admits an rr-matrix rr, then we will abbreviate this by writing GrG_{r} and 𝔤r\mathfrak{g}_{r}. If [r,r]=[c,c][r,r]=-[c,c] and cc defines a perfect pairing, then we say that rr is factorisable, and we will say that the Lie bialgebra 𝔤r\mathfrak{g}_{r} and any corresponding Poisson algebraic group GrG_{r} are factorisable.

Notation 4.14.

We denote the torus component of such cc by c𝔱c_{\mathfrak{t}}. Let LgL_{g} and RgR_{g} denote the translations on GG of left and right multiplication by gg. Given an element r𝔤𝔤r\in\mathfrak{g}\otimes\mathfrak{g}, we then write rR:=rR,R:=(dRgdRg)(r)r^{R}:=r^{R,R}:=(\mathrm{d}R_{g}\otimes\mathrm{d}R_{g})(r), rR,L:=(dRgdLg)(r)r^{R,L}:=(\mathrm{d}R_{g}\otimes\mathrm{d}L_{g})(r), etc. Then set rad:=rR,R+rL,LrR,LrL,Rr^{\mathrm{ad}}:=r^{R,R}+r^{L,L}-r^{R,L}-r^{L,R}.

Semenov-Tian-Shansky used Proposition 4.12 to geometrically prove over \mathbb{C} the following

Theorem 4.15 ([STS85, p. 1247]).

Given a factorisable Poisson algebraic group GrG_{r} over a scheme where 2 is invertible, let cc in Sym2(𝔤)𝔤\mathrm{Sym}^{2}(\mathfrak{g})^{\mathfrak{g}} be such that [r,r]=[c,c][r,r]=-[c,c], let GG_{*} denote the underlying scheme of GG but now equipped with the bivector

12(radcL,R+cR,L)=(r+R,Rr+L,R)(rR,LrL,L).\frac{1}{2}(r^{\mathrm{ad}}-c^{L,R}+c^{R,L})=(r_{+}^{R,R}-r_{+}^{L,R})-(r_{-}^{R,L}-r_{-}^{L,L}). (4.2)

This yields a Poisson structure on GG, and the right conjugation map Gr×GGG_{r}\times G_{*}\rightarrow G_{*} is Poisson.

This can also be proven directly, without using factorisability.

Definition 4.16.

This is called the (right) Semenov-Tian-Shansky bracket on GG.

Notation 4.17.

In order to obtain more such brackets in low characteristic, we slightly enlarge 𝔱\mathfrak{t} to by adding the dual basis {ωˇi}i=1rk\{\check{\omega}_{i}\}_{i=1}^{\mathrm{rk}} to the simple roots and denote the result by 𝔱sc\mathfrak{t}_{\mathrm{sc}}. We then assume that c𝔱c_{\mathfrak{t}} lies in 𝔱sc𝔱\mathfrak{t}_{\mathrm{sc}}\otimes\mathfrak{t}:

Proposition 4.18.

If 𝔤\mathfrak{g} is defined over an arbitrary ring then such c𝔱c_{\mathfrak{t}} and r±r_{\pm} might not lie in 𝔤\mathfrak{g}, but the corresponding Semenov-Tian-Shansky bracket still yields integral formulas.

Proof.

We only prove the second part. In the left-trivialisation of the tangent bundle of GG, the right-hand-side of equation (4.2) is given at any point gg in GG by

((Adgid)Adg)(r+)((Adgid)id)(r),\bigl{(}(\mathrm{Ad}_{g}-\mathrm{id})\otimes\mathrm{Ad}_{g}\bigl{)}(r_{+})-\bigl{(}(\mathrm{Ad}_{g}-\mathrm{id})\otimes\mathrm{id}\bigl{)}(r_{-}),

whose torus component is

((Adgid)(Adg+id))(c𝔱/2).\bigl{(}(\mathrm{Ad}_{g}-\mathrm{id})\otimes(\mathrm{Ad}_{g}+\mathrm{id})\bigl{)}(c_{\mathfrak{t}}/2).

We may decompose gg into a product of root subgroups, and induct on the length of such an expression. By projecting xx to root spaces, it suffices to prove the claim for elements of the form (ωˇiti/2)(\check{\omega}_{i}\otimes t_{i}/2). For gg an element of a root subgroups, it follows from

Adexpβ(y)(x)=x+β(x)y,\mathrm{Ad}_{\exp_{\beta}(y)}(x)=x+\beta(x)y,

where expβ:𝔤βNβ\exp_{\beta}:\mathfrak{g}_{\beta}\rightarrow N_{\beta} is the exponential map and yy lies in 𝔤β\mathfrak{g}_{\beta}. The claim now follows by induction on the length of the decomposition of gg: if g=gg′′g=g^{\prime}g^{\prime\prime} with g′′g^{\prime\prime} lying in a root subgroup, then

Adg(x)±x=Adg′′(Adg(x)±x)(Adg′′(x)x)Adg′′(𝔤)+𝔤=𝔤.\mathrm{Ad}_{g}(x)\pm x=\mathrm{Ad}_{g^{\prime\prime}}\bigl{(}\mathrm{Ad}_{g^{\prime}}(x)\pm x\bigr{)}\mp\bigl{(}\mathrm{Ad}_{g^{\prime\prime}}(x)-x\bigr{)}\in\mathrm{Ad}_{g^{\prime\prime}}(\mathfrak{g})+\mathfrak{g}=\mathfrak{g}.\qed
Example 4.19.

Let 𝔤\mathfrak{g} be a reductive Lie algebra over a field. Extend Chevalley generators to a basis {eβ,fβ}β+\{e_{\beta},f_{\beta}\}_{\beta\in\mathfrak{R}_{+}} and {ti}i=1rk\{t_{i}\}_{i=1}^{\mathrm{rk}} and then denote by {ωˇi}i=1rk\{\check{\omega}_{i}\}_{i=1}^{\mathrm{rk}} the dual basis in 𝔱\mathfrak{t} to the simple roots. If we set

r:=β+dβfβeβ,c:=β+dβeβfβ+β+dβfβeβ+i=1rkωˇiti,r:=\sum_{\beta\in\mathfrak{R}_{+}}d_{\beta}f_{\beta}\wedge e_{\beta},\qquad c:=\sum_{\beta\in\mathfrak{R}_{+}}d_{\beta}e_{\beta}\otimes f_{\beta}+\sum_{\beta\in\mathfrak{R}_{+}}d_{\beta}f_{\beta}\otimes e_{\beta}+\sum_{i=1}^{\mathrm{rk}}\check{\omega}_{i}\otimes t_{i},

where dβ{1,2,3}d_{\beta}\in\{1,2,3\} are the usual symmetrisers with dβ=1d_{\beta}=1 for long roots, then this Casimir element yields [r,r]=[c,c][r,r]=-[c,c].

Notation 4.20.

Given a factorisable rr-matrix rr and cc in Sym2(𝔤)𝔤\mathrm{Sym}^{2}(\mathfrak{g})^{\mathfrak{g}} such that [r,r]=[c,c][r,r]=-[c,c], we denote the inner product corresponding to cc by (,)r(\cdot,\cdot)_{r}, so that the orthogonal complement of a submodule 𝔥\mathfrak{h} of 𝔤\mathfrak{g} is given by 𝔥:=c(𝔥0)\mathfrak{h}^{\bot}:=c(\mathfrak{h}^{0}). We furthermore write 𝔟±r:=r±(𝔤)\mathfrak{b}_{\pm}^{r}:=r_{\pm}(\mathfrak{g}^{*}) and 𝔪±r:=r±(kerr)\mathfrak{m}_{\pm}^{r}:=r_{\pm}(\ker\,r_{\mp}).

Proposition 4.21.

Let 𝔤r\mathfrak{g}_{r} be a factorisable Lie bialgebra.

  1. (i)

    [BD82] The annihilator of 𝔟±r\mathfrak{b}_{\pm}^{r} in 𝔤\mathfrak{g}^{*} is kerr\mathrm{ker}\,r_{\mp}.

  2. (ii)

    Let 𝔥\mathfrak{h} be a subspace of 𝔤\mathfrak{g} containing 𝔟±r\mathfrak{b}_{\pm}^{r}. Then 𝔥=r±(𝔥0)𝔪±\mathfrak{h}^{\bot}=r_{\pm}(\mathfrak{h}^{0})\subseteq\mathfrak{m}_{\pm}, with equality if and only if 𝔥0=(𝔟±r)0\mathfrak{h}^{0}=(\mathfrak{b}_{\pm}^{r})^{0}.

    In particular, we have

    ιr(𝔥0)=(𝔥,0)(resp. ιr(𝔥0)=(0,𝔥)).\iota_{r}(\mathfrak{h}^{0})=(\mathfrak{h}^{\bot},0)\qquad\bigl{(}\textrm{resp. }\iota_{r}(\mathfrak{h}^{0})=(0,\mathfrak{h}^{\bot})\bigr{)}.
Proof.

(i): Any element of 𝔟±r\mathfrak{b}_{\pm}^{r} is of the form r±xr_{\pm}x for some x𝔤x\in\mathfrak{g}^{*}. For y𝔤y\in\mathfrak{g}^{*} we then have

y,r±x=r±21y,x=ry,x.\langle y,r_{\pm}x\rangle=\langle r_{\pm}^{21}y,x\rangle=\langle-r_{\mp}y,x\rangle.

Since xx can be chosen arbitrarily, the claim follows.

(ii): From (i) it follows that 𝔥0(𝔟±r)0=kerr\mathfrak{h}^{0}\subseteq(\mathfrak{b}_{\pm}^{r})^{0}=\mathrm{ker}\,r_{\mp}, so

𝔥=c(𝔥0)(r+r)(kerr)=±r±(kerr)=:𝔪±.\mathfrak{h}^{\bot}=c(\mathfrak{h}^{0})\subseteq(r_{+}-r_{-})(\mathrm{ker}\,r_{\mp})=\pm r_{\pm}(\mathrm{ker}\,r_{\mp})=:\mathfrak{m}_{\pm}.

The claim on equality follows from nondegeneracy of cc. For 𝔥\mathfrak{h} containing 𝔟+r\mathfrak{b}_{+}^{r} we now find

ιr(𝔥0)=(r+,r)(𝔥0)=(r+(𝔥0),0)=(𝔥,0).\iota_{r}(\mathfrak{h}^{0})=(r_{+},r_{-})(\mathfrak{h}^{0})=\bigl{(}r_{+}(\mathfrak{h}^{0}),0\bigr{)}=(\mathfrak{h}^{\bot},0).\qed
Definition 4.22.

A Belavin-Drinfeld triple is a triple 𝔗=(𝔗0,𝔗1,τ)\mathfrak{T}=(\mathfrak{T}_{0},\mathfrak{T}_{1},\tau) where 𝔗0,𝔗1\mathfrak{T}_{0},\mathfrak{T}_{1} are sets of simple roots and τ:𝔗0𝔗1\tau:\mathfrak{T}_{0}\rightarrow\mathfrak{T}_{1} is a bijection such that

  • (τ(α),τ(α~))=(α,α~)\bigl{(}\tau(\alpha),\tau(\tilde{\alpha})\bigr{)}=(\alpha,\tilde{\alpha}) for any pair of simple roots α,α~𝔗0\alpha,\tilde{\alpha}\in\mathfrak{T}_{0}, and

  • τ\tau is nilpotent: for any α𝔗0\alpha\in\mathfrak{T}_{0} there exists m1m\in\mathbb{N}_{1} such that τm(α)𝔗1\𝔗0\tau^{m}(\alpha)\in\mathfrak{T}_{1}\backslash\mathfrak{T}_{0}.

Such a triple gives the set of positive roots a partial ordering: for positive roots β,β\beta,\beta^{\prime} we set β<β\beta<\beta^{\prime} if β0𝔗0\beta\in\mathbb{N}_{0}\mathfrak{T}_{0}, β0𝔗1\beta^{\prime}\in\mathbb{N}_{0}\mathfrak{T}_{1} and τm(β)=β\tau^{m}(\beta)=\beta^{\prime} for some m1m\in\mathbb{N}_{1}.

Theorem 4.23 ([BD82]).

Let 𝔤r\mathfrak{g}_{r} be a factorisable reductive Lie bialgebra over a field of characteristic 0 (or sufficiently large). Then there exists a Cartan decomposition and a Belavin-Drinfeld triple 𝔗=(𝔗0,𝔗1,τ)\mathfrak{T}=(\mathfrak{T}_{0},\mathfrak{T}_{1},\tau) such that

r+\displaystyle r_{+} =12(r0+c𝔱)+β+dβfβeβ+β<βdβfβeβ,\displaystyle=\frac{1}{2}(r_{0}+c_{\mathfrak{t}})+\sum_{\beta\in\mathfrak{R}_{+}}d_{\beta}f_{\beta}\otimes e_{\beta}+\sum_{\beta<\beta^{\prime}}d_{\beta}f_{\beta}\wedge e_{\beta^{\prime}}, (4.3)

where fβf_{\beta} and eβe_{\beta} are root vectors with weight β-\beta and β\beta respectively, are normalised by dβ(fβ,eβ)=1d_{\beta}(f_{\beta},e_{\beta})=1 for an invariant bilinear form corresponding to some element cSym2(𝔤)𝔤c\in\mathrm{Sym}^{2}(\mathfrak{g})^{\mathfrak{g}}, whilst the element r0𝔱𝔱r_{0}\in\mathfrak{t}\wedge\mathfrak{t} satisfies

(τ(α)1)(r0+c𝔱)+(1α)(r0+c𝔱)=0,α𝔗0,\bigl{(}\tau(\alpha)\otimes 1\bigr{)}(r_{0}+c_{\mathfrak{t}})+(1\otimes\alpha)(r_{0}+c_{\mathfrak{t}})=0,\qquad\forall\alpha\in\mathfrak{T}_{0}, (4.4)

where c𝔱c_{\mathfrak{t}} is the image of the 𝔱\mathfrak{t}-component of cc.

Furthermore, such solutions r0r_{0} form a torsor for the k𝔗(k𝔗1)/2k_{\mathfrak{T}}(k_{\mathfrak{T}}-1)/2-dimensional vector space 𝔱𝔗𝔱𝔗\mathfrak{t}_{\mathfrak{T}}\wedge\mathfrak{t}_{\mathfrak{T}}, where k𝔗:=rank(𝔤)|𝔗0|k_{\mathfrak{T}}:=\mathrm{rank}(\mathfrak{g})-|\mathfrak{T}_{0}| and

𝔱𝔗:={t𝔱:β(t)=β(t) for all β<β}.\mathfrak{t}_{\mathfrak{T}}:=\{t\in\mathfrak{t}:\beta(t)=\beta^{\prime}(t)\text{ for all }\beta<\beta^{\prime}\}. (4.5)

Explicitly, we then have

r\displaystyle r_{-} =12(r0c𝔱)β+dβeβfβ+β<βdβfβeβ.\displaystyle=\frac{1}{2}(r_{0}-c_{\mathfrak{t}})-\sum_{\beta\in\mathfrak{R}_{+}}d_{\beta}e_{\beta}\otimes f_{\beta}+\sum_{\beta<\beta^{\prime}}d_{\beta}f_{\beta}\wedge e_{\beta^{\prime}}. (4.6)
Proposition 4.24.

Let GrG_{r} be a reductive factorisable Poisson algebraic group and let HGrH\subseteq G_{r} be a connected closed subgroup with Lie algebra 𝔥\mathfrak{h}.

  1. (i)

    If 𝔥\mathfrak{h} contains 𝔟+r\mathfrak{b}_{+}^{r} or 𝔟r\mathfrak{b}_{-}^{r} then HH is Poisson.

  2. (ii)

    Now suppose that the rr-matrix comes from a Belavin-Drinfeld triple 𝔗\mathfrak{T}, and that HH is a product of root subgroups corresponding to a subset of positive (resp. negative) roots 𝔑\mathfrak{N} of the form w\mathfrak{R}_{w}. If 𝔗𝔑c\mathfrak{T}\subseteq\mathfrak{N}^{c} then HH is coisotropic.

Proof.

(i): Let’s assume that 𝔥\mathfrak{h} contains 𝔟+r\mathfrak{b}_{+}^{r}, then by Proposition 4.21(ii) we have ιr(𝔥0)=(𝔥,0)\iota_{r}(\mathfrak{h}^{0})=(\mathfrak{h}^{\bot},0). Thus, in order to prove that 𝔥0\mathfrak{h}^{0} is an ideal of 𝔤\mathfrak{g}^{*}, it suffices to show that [𝔟+r,𝔥]𝔥[\mathfrak{b}_{+}^{r},\mathfrak{h}^{\bot}]\subseteq\mathfrak{h}^{\bot}. Given x𝔟+r𝔥x\in\mathfrak{b}_{+}^{r}\subseteq\mathfrak{h} and y𝔥y\in\mathfrak{h}^{\bot} and z𝔥z\in\mathfrak{h}, invariance of (,)r(\cdot,\cdot)_{r} yields

([x,y],z)r=(y,[z,x])r=0([x,y],z)_{r}=(y,[z,x])_{r}=0

so that [x,y]𝔥[x,y]\in\mathfrak{h}^{\bot}.

(ii): Let {eβ,fβ}β+\{e_{\beta}^{*},f_{\beta}^{*}\}_{\beta\in\mathfrak{R}_{+}} be the basis of 𝔫+𝔫𝔤\mathfrak{n}_{+}^{*}\oplus\mathfrak{n}_{-}^{*}\subset\mathfrak{g}^{*} dual to the usual basis {eβ,fβ}β+\{e_{\beta},f_{\beta}\}_{\beta\in\mathfrak{R}_{+}} of 𝔫+𝔫\mathfrak{n}_{+}\oplus\mathfrak{n}_{-}, so eβe_{\beta}^{*} is the element of 𝔤\mathfrak{g}^{*} vanishing on 𝔱\mathfrak{t} and all root spaces other than 𝔫β\mathfrak{n}_{\beta}, where it is given on its basis vector by eβ(eβ)=1e_{\beta}^{*}(e_{\beta})=1, etc. Then

ιr(eβ)=dβ(i>0fτi(β),i0fτi(β)),ιr(fβ)=dβ(i0eτi(β),i>0eτi(β))\iota_{r}(e_{\beta}^{*})=-d_{\beta}(\sum_{i>0}f_{\tau^{-i}(\beta)},\sum_{i\geq 0}f_{\tau^{-i}(\beta)}),\qquad\iota_{r}(f_{\beta}^{*})=d_{\beta}(\sum_{i\geq 0}e_{\tau^{i}(\beta)},\sum_{i>0}e_{\tau^{i}(\beta)}) (4.7)

where we set eτi+1(β)=0e_{\tau^{i+1}(\beta)}=0 (resp. fτi1(β)=0f_{\tau^{-i-1}(\beta)}=0) when τi(β)\tau^{i}(\beta) is not in the span of 𝔗0\mathfrak{T}_{0} (resp. τi(β)\tau^{-i}(\beta) not in the span of 𝔗1\mathfrak{T}_{1}). In particular, if β\beta is not in 𝔑c\mathfrak{N}^{c} then it is not in 𝔗0\mathfrak{T}_{0} or 𝔗1\mathfrak{T}_{1} and we find

ιr(eβ)=dβ(0,fβ),ιr(fβ)=dβ(eβ,0).\iota_{r}(e_{\beta}^{*})=-d_{\beta}(0,f_{\beta}),\qquad\iota_{r}(f_{\beta}^{*})=d_{\beta}(e_{\beta},0). (4.8)

Since by assumption 𝔥\mathfrak{h} is a sum of root subspaces, its dual is spanned by a subset of {eβ,fβ}β+\{e_{\beta}^{*},f_{\beta}^{*}\}_{\beta\in\mathfrak{R}_{+}} plus a basis for 𝔱\mathfrak{t}^{*}. The adjoint action of elements of ιr(𝔱)\iota_{r}(\mathfrak{t}^{*}) on elements in (4.7) is through scalars due to (4.5), so we can safely ignore them.

Let’s do the positive case: since 𝔗𝔑c\mathfrak{T}\subseteq\mathfrak{N}^{c} and the set of roots 𝔑c\mathfrak{N}^{c} is convex [Pap94], it follows from (4.7) that bracket of ιr(eβ)\iota_{r}(e_{\beta}^{*}) and ιr(eβ)\iota_{r}(e_{\beta^{\prime}}^{*}) for β\beta in 𝔑c\mathfrak{N}^{c} corresponds to root vectors lying in (4.7). As 𝔤\mathfrak{g} is closed under bracketing, (4.8) implies that it must be a linear combination of elements ιr(eβ′′)\iota_{r}(e_{\beta^{\prime\prime}}^{*}) with β′′\beta^{\prime\prime} in 𝔑c\mathfrak{N}^{c}: there are simply no other elements in ιr(𝔤)\iota_{r}(\mathfrak{g}^{*}) projecting to the right weight spaces.

For ιr(eβ)\iota_{r}(e_{\beta}^{*}) and ιr(fβ)\iota_{r}(f_{\beta^{\prime}}^{*}) with β𝔑c\beta\in\mathfrak{N}^{c} and β+\beta^{\prime}\in\mathfrak{R}_{+}, note that only the parts lying in 𝔗\mathfrak{T} can bracket nontrivially. As 𝔗\mathfrak{T} forms a standard parabolic subsystem it similarly follows that the bracket is a linear combination of elements ιr(eβ′′)\iota_{r}(e_{\beta^{\prime\prime}}^{*}) with β′′\beta^{\prime\prime} in 𝔑c\mathfrak{N}^{c} and ιr(fβ′′′)\iota_{r}(f_{\beta^{\prime\prime\prime}}^{*}) with β′′′\beta^{\prime\prime\prime} arbitrary. ∎

Restricting the left or right multiplication of GG on itself to HH as in (i), we thus obtain Poisson structures on the GIT quotients H\\GH\backslash\!\backslash G and G//HG/\!/H. In the particular case where rr is the standard rr-matrix (so that N±r=N±N_{\pm}^{r}=N_{\pm} and B±r=B±B_{\pm}^{r}=B_{\pm}) and HH is a parabolic subgroup, the corresponding Poisson structure has been extensively studied (e.g. [GY09]).

4.3 Reducing the Semenov-Tian-Shansky bracket

Lemma 4.25.

Let GG be a Poisson group scheme with an action on a Poisson scheme XX. Let HGH\subseteq G be a closed subgroup preserving a smooth closed subscheme ι:ZX\iota:Z\hookrightarrow X such that (ι)1(𝒪ZH)(\iota^{\sharp})^{-1}(\mathcal{O}_{Z}^{H}) is a sheaf of Poisson subalgebras of ι1(𝒪X)\iota^{-1}(\mathcal{O}_{X}). Then the following are equivalent:

  1. (i)

    The ideal sheaf Z\mathcal{I}_{Z} is a subsheaf of Poisson ideals of (ι)1(𝒪ZH)(\iota^{\sharp})^{-1}(\mathcal{O}_{Z}^{H}); in other words, the Poisson bracket on (ι)1(𝒪ZH)(\iota^{\sharp})^{-1}(\mathcal{O}_{Z}^{H}) reduces to 𝒪ZH\mathcal{O}_{Z}^{H}.

  2. (ii)

    For any function ff in (ι)1(𝒪ZH)(\iota^{\sharp})^{-1}(\mathcal{O}_{Z}^{H}), its Hamiltonian vector field lies in TZT_{Z}.

Proof.

(i) \Rightarrow (ii): Pick a function ff in (ι)1(𝒪ZH)(\iota^{\sharp})^{-1}(\mathcal{O}_{Z}^{H}) and a covector α\alpha at zz which is annihilated by all tangent vectors in TzZT_{z}Z. Lifting α\alpha to a one-form in a neighbourhood of zz, by the conormal exact sequence we can find a function ff^{\prime} in Z\mathcal{I}_{Z} such that df|z=α\mathrm{d}f^{\prime}|_{z}=\alpha and f|Z=0f^{\prime}|_{Z}=0. Then the hypothesis yields that

α(Hamf)={f,f}|z={f|Z,0}=0,\alpha\bigl{(}\mathrm{Ham}_{f}\bigr{)}=\{f,f^{\prime}\}|_{z}=\{f|_{Z},0\}=0,

which means that Hamf\mathrm{Ham}_{f} is annihilated by the covectors annihilated by TzZT_{z}Z, which by smoothness implies that it lies in the tangent space TzZT_{z}Z.

(ii) \Rightarrow (i): Let ff be a function in (ι)1(𝒪ZH)(\iota^{\sharp})^{-1}(\mathcal{O}_{Z}^{H}) and ff^{\prime} in Z\mathcal{I}_{Z}, then from the inclusion HamfTZ\mathrm{Ham}_{f}\in T_{Z} it follows that

{f,f}|Z=df(Hamf)=d(f|Z)(Hamf)=0,\{f,f^{\prime}\}|_{Z}=\mathrm{d}f^{\prime}(\mathrm{Ham}_{f})=\mathrm{d}(f^{\prime}|_{Z})(\mathrm{Ham}_{f})=0,

implying that {f,f}\{f,f^{\prime}\} lies in Z\mathcal{I}_{Z}. ∎

Corollary 4.26.

Equip GG with the right Semenov-Tian-Shansky bracket coming from a factorisable rr-matrix with 𝔗𝔏\mathfrak{T}\subseteq\mathfrak{L}. Consider the right conjugation action of NGN\subset G on GG_{*} and its restriction to the closed subscheme ι:Nw˙LNG\iota:N\dot{w}LN\hookrightarrow G_{*}. The subgroup NN is coisotropic and (ι)1(𝒪Nw˙LNN)(\iota^{\sharp})^{-1}(\mathcal{O}_{N\dot{w}LN}^{N}) is a subsheaf of ι1(𝒪G)\iota^{-1}(\mathcal{O}_{G_{*}}) of Poisson subalgebras.

Proof.

This follows by combining Proposition 4.24 with Proposition 4.9. ∎

Notation 4.27.

For any function ff in 𝒪G\mathcal{O}_{G} and element gg in GG, we define dfg𝔤\mathrm{d}f^{g}\in\mathfrak{g}^{*} by evaluating on a tangent vector in 𝔤\mathfrak{g} as follows:

dfg():=df(Rg()Lg())𝔤,\mathrm{d}f^{g}(\cdot):=\mathrm{d}f\bigl{(}R_{g}(\cdot)-L_{g}(\cdot)\bigr{)}\in\mathfrak{g}^{*},

so in the left trivialisation we have dfg()=df(Adg()id())\mathrm{d}f^{g}(\cdot)=\mathrm{d}f\bigl{(}\mathrm{Ad}_{g}(\cdot)-\mathrm{id}(\cdot)\bigr{)}. The tangent space to Nw˙LN=NLw˙NN\dot{w}LN=NL\dot{w}N at one of its elements gg is given by

Tg:=Adg(𝔫)+𝔩𝔫=Adg(𝔫𝔩)+𝔫𝔤,T_{g}:=\mathrm{Ad}_{g}(\mathfrak{n})+\mathfrak{l}\oplus\mathfrak{n}=\mathrm{Ad}_{g}(\mathfrak{n}\oplus\mathfrak{l})+\mathfrak{n}\subseteq\mathfrak{g}, (4.9)

where Adg()\mathrm{Ad}_{g}(\cdot) still denotes the right adjoint action.

Lemma 4.28.

Assume that 𝔏\mathfrak{L} is a standard parabolic subroot system. Then

Tg𝔱=𝔩𝔱.T_{g}\cap\mathfrak{t}=\mathfrak{l}\cap\mathfrak{t}.
Proof.

Let g=nlw˙n~g=nl\dot{w}\tilde{n}. Since 𝔑\mathfrak{N} is convex, we have Adn(𝔫)𝔫\mathrm{Ad}_{n}(\mathfrak{n})\subseteq\mathfrak{n}. From the assumption on 𝔏\mathfrak{L} it follows that

Adl(𝔫)β+\𝔏𝔤β.\mathrm{Ad}_{l}(\mathfrak{n})\subseteq\bigoplus_{\beta\in\mathfrak{R}_{+}\backslash\mathfrak{L}}\mathfrak{g}_{\beta}.

As 𝔏\mathfrak{L} is preserved by ww, it now follows that

Adnlw˙(𝔫)Adw˙(Adl(𝔫))Adw˙(β+\𝔏𝔤β)β\𝔏𝔤β.\mathrm{Ad}_{nl\dot{w}}(\mathfrak{n})\subseteq\mathrm{Ad}_{\dot{w}}\bigl{(}\mathrm{Ad}_{l}(\mathfrak{n})\bigr{)}\subseteq\mathrm{Ad}_{\dot{w}}\bigl{(}\bigoplus_{\beta\in\mathfrak{R}_{+}\backslash\mathfrak{L}}\mathfrak{g}_{\beta}\bigr{)}\subseteq\bigoplus_{\beta\in\mathfrak{R}\backslash\mathfrak{L}}\mathfrak{g}_{\beta}.

Now let xx be an arbitrary element in the right-hand-side. If β\beta is of minimal height among the roots in \𝔏\mathfrak{R}\backslash\mathfrak{L} such that the projection of xx to 𝔤β\mathfrak{g}_{\beta} is nonzero, then the projection of Adn~(x)\mathrm{Ad}_{\tilde{n}}(x) to 𝔤β\mathfrak{g}_{\beta} is still nonzero. Thus if β\beta is negative, Adn~(x)\mathrm{Ad}_{\tilde{n}}(x) does not lie in 𝔱+𝔩𝔫\mathfrak{t}+\mathfrak{l}\oplus\mathfrak{n}; if β\beta is positive then Adn~(x)\mathrm{Ad}_{\tilde{n}}(x) still lies in 𝔫+\mathfrak{n}_{+}. Hence

Tg𝔱=(Adg(𝔫)+𝔩𝔫)𝔱=(𝔩𝔫)𝔱=𝔩𝔱.T_{g}\cap\mathfrak{t}=\bigl{(}\mathrm{Ad}_{g}(\mathfrak{n})+\mathfrak{l}\oplus\mathfrak{n}\bigr{)}\cap\mathfrak{t}=(\mathfrak{l}\oplus\mathfrak{n})\cap\mathfrak{t}=\mathfrak{l}\cap\mathfrak{t}.\qed
Proposition 4.29.

Fix an ordering β1,,βl\beta_{1},\ldots,\beta_{l} of the roots of w\mathfrak{R}_{w} by height, and fix the element g:=pβ1(1)pβl(1)g^{\prime}:=p_{\beta_{1}}(1)\cdots p_{\beta_{l}}(1). Then

𝔱sc𝔫w,tproj𝔫w(Adg(t)t)\mathfrak{t}_{\mathrm{sc}}\longrightarrow\mathfrak{n}_{w},\qquad t\longmapsto\mathrm{proj}_{\mathfrak{n}_{w}}\bigl{(}\mathrm{Ad}_{g}(t)-t\bigr{)}

is injective.

Proof.

By construction of 𝔱sc\mathfrak{t}_{\mathrm{sc}}, we can recover the coordinates of tt in the standard basis of 𝔱sc\mathfrak{t}_{\mathrm{sc}} through studying tα(t)t\mapsto\alpha(t) as α\alpha ranges over the simple root. By [Mal21, Corollary 3.13], these values can be obtained from the roots in w\mathfrak{R}_{w}. For β\beta in w\mathfrak{R}_{w}, let eβe_{\beta} denote the element of 𝔤β\mathfrak{g}_{\beta} exponentiating to pβ(1)p_{\beta}(1). We write O(βi+1)O(\beta_{i+1}) to mean a polynomial in the eβje_{\beta_{j}} with ji+1j\geq i+1. It follows from

Adg(t)t\displaystyle\mathrm{Ad}_{g^{\prime}}(t)-t =Adpβ2(1)pβl(1)(t+β1(t)eβ1)\displaystyle=\mathrm{Ad}_{p_{\beta_{2}}(1)\cdots p_{\beta_{l}}(1)}\bigl{(}t+\beta_{1}(t)e_{\beta_{1}}\bigr{)}
=Adpβ3(1)pβl(1)(t+β1(t)eβ1+β2(t)eβ2)+β1(t)Adpβ2(1)(eβ1)t\displaystyle=\mathrm{Ad}_{p_{\beta_{3}}(1)\cdots p_{\beta_{l}}(1)}\bigl{(}t+\beta_{1}(t)e_{\beta_{1}}+\beta_{2}(t)e_{\beta_{2}})+\beta_{1}(t)\mathrm{Ad}_{p_{\beta_{2}}(1)}\bigl{(}e_{\beta_{1}}\bigr{)}-t
=Adpβ3(1)pβl(1)(t+β1(t)eβ1+β2(t)eβ2)+β1(t)O(β2)t\displaystyle=\mathrm{Ad}_{p_{\beta_{3}}(1)\cdots p_{\beta_{l}}(1)}\bigl{(}t+\beta_{1}(t)e_{\beta_{1}}+\beta_{2}(t)e_{\beta_{2}})+\beta_{1}(t)O(\beta_{2})-t
=i=1l(βi(t)eβi+O(βi+1))\displaystyle=\sum_{i=1}^{l}\bigl{(}\beta_{i}(t)e_{\beta_{i}}+O(\beta_{i+1})\bigr{)}

that the values βi(t)\beta_{i}(t) can be recovered inductively. ∎

Corollary 4.30.

Now set g:=w˙gg:=\dot{w}g^{\prime}. The map

(ι)1(𝒪Nw˙LNN)𝔱sc,f(tdfg(t))(\iota^{\sharp})^{-1}(\mathcal{O}_{N\dot{w}LN}^{N})\longrightarrow\mathfrak{t}_{\mathrm{sc}}^{*},\qquad f\longmapsto\bigl{(}t\mapsto\mathrm{d}f^{g}(t)\bigr{)}

is surjective.

Proof.

By the previous statement, we can recover w1(t)w^{-1}(t) and hence tt from the projection of

Adg(t)w1(t)=Adg(w1(t))w1(t)\mathrm{Ad}_{g}\bigl{(}t\bigr{)}-w^{-1}(t)=\mathrm{Ad}_{g^{\prime}}\bigl{(}w^{-1}(t)\bigr{)}-w^{-1}(t)

to 𝔫w\mathfrak{n}_{w}, and then the same is true for

tg:=(Adgid)(t)=(Adg(t)w1(t))+(w1(t)t)t^{g}:=(\mathrm{Ad}_{g}-\mathrm{id})(t)=\bigl{(}\mathrm{Ad}_{g}\bigl{(}t\bigr{)}-w^{-1}(t)\bigr{)}+\bigr{(}w^{-1}(t)-t\bigr{)}

since we are adding an element of 𝔱\mathfrak{t}. Thus

𝒪w˙Nw𝔱sc\mathcal{O}_{\dot{w}N_{w}}\longrightarrow\mathfrak{t}_{\mathrm{sc}}^{*}

is surjective. As w˙LNw=Lw˙NwL×w˙Nw\dot{w}LN_{w}=L\dot{w}N_{w}\simeq L\times\dot{w}N_{w}, the the cross section isomorphism

𝒪Nw˙LN𝒪N𝒪w˙LNw𝒪N𝒪L𝒪w˙Nw\mathcal{O}_{N\dot{w}LN}\overset{\sim}{\longrightarrow}\mathcal{O}_{N}\otimes\mathcal{O}_{\dot{w}LN_{w}}\overset{\sim}{\longrightarrow}\mathcal{O}_{N}\otimes\mathcal{O}_{L}\otimes\mathcal{O}_{\dot{w}N_{w}}

now yields the claim as elements of the form 11f1\otimes 1\otimes f^{\prime} pull back to NN-invariants in 𝒪Nw˙LN\mathcal{O}_{N\dot{w}LN} satisfying

dfg(t)=df(tg)=d(1f)(tg).\mathrm{d}f^{g}(t)=\mathrm{d}f(t^{g})=\mathrm{d}(1\cdot f^{\prime})(t^{g}).

Example 4.31.

Consider the usual group

G=SL2={[abcd]:adbc=1}G=\mathrm{SL}_{2}=\left\{\begin{bmatrix}a&b\\ c&d\end{bmatrix}:ad-bc=1\right\}

of type 𝖠1\mathsf{A}_{1} and let

w˙=[0110]\dot{w}=\begin{bmatrix}0&1\\ -1&0\end{bmatrix}

be the usual lift of its Coxeter element. Let tt be the usual diagonal element of its Lie algebra and

g=w˙pα(1)=[0111],g=\dot{w}p_{\alpha}(1)=\begin{bmatrix}0&1\\ -1&-1\end{bmatrix},

then

RgtLgt=[0220].R_{g}t-L_{g}t=\begin{bmatrix}0&2\\ 2&0\end{bmatrix}.

The matrix coordinate cc is invariant under the conjugation action of N,N, and dcg(t)=2\mathrm{d}c^{g}(t)=2.

The following lemma was obtained by dissecting Sevostyanov’s proof [Sev11, Theorem 5.2]:

Notation 4.32.

Given a function ff in 𝒪G\mathcal{O}_{G} defined at a closed point gg of GG, we write

df±g:=r±(dfg)𝔤\mathrm{d}f_{\pm}^{g}:=r_{\pm}(\mathrm{d}f^{g})\in\mathfrak{g}

and then decompose these elements along the Cartan decomposition 𝔤=𝔱(𝔫+𝔫)\mathfrak{g}=\mathfrak{t}\oplus(\mathfrak{n}_{+}\oplus\mathfrak{n}_{-}) as

df±g=df±0g+df±g,df±0g𝔱,df±g𝔫+𝔫.\mathrm{d}f_{\pm}^{g}=\mathrm{d}f_{\pm 0}^{g}+\mathrm{d}f_{\pm\emptyset}^{g},\qquad\mathrm{d}f_{\pm 0}^{g}\in\mathfrak{t},\quad\mathrm{d}f_{\pm\emptyset}^{g}\in\mathfrak{n}_{+}\oplus\mathfrak{n}_{-}.

From (4.3) and (4.6) one then sees that

df±0g=dfg(r0±c𝔱)\mathrm{d}f_{\pm 0}^{g}=\mathrm{d}f^{g}(r_{0}\pm c_{\mathfrak{t}})

and similarly for df±g\mathrm{d}f_{\pm\emptyset}^{g}.

From the right-hand-side of equation (4.2), it follows that Hamiltonian vector field of a function ff in 𝒪G\mathcal{O}_{G_{*}} is given at a point gg in this trivialisation by

Hamf(g)=Adg(r+(dfg))r(dfg).\mathrm{Ham}_{f}(g)=\mathrm{Ad}_{g}\bigl{(}r_{+}(\mathrm{d}f^{g})\bigr{)}-r_{-}(\mathrm{d}f^{g}).
Lemma 4.33.

For any function ff in (ι)1(𝒪Nw˙LNN)(\iota^{\sharp})^{-1}(\mathcal{O}_{N\dot{w}LN}^{N}) defined at a point gg of Nw˙LNN\dot{w}LN, we have

Hamf(g)Adw˙(df+0g)df0gmodTg.\mathrm{Ham}_{f}(g)\equiv\mathrm{Ad}_{\dot{w}}(\mathrm{d}f_{+0}^{g})-\mathrm{d}f_{-0}^{g}\mod T_{g}.
Proof.

The slicing assumption 𝔏𝔑=+\mathfrak{L}\sqcup\mathfrak{N}=\mathfrak{R}_{+} implies that 𝔩+𝔱𝔫+=𝔱+𝔩𝔫\mathfrak{l}+\mathfrak{t}\oplus\mathfrak{n}_{+}=\mathfrak{t}+\mathfrak{l}\oplus\mathfrak{n}, so from 𝔗0𝔏\mathfrak{T}_{0}\subseteq\mathfrak{L} we deduce that

𝔟+r𝔩+𝔱𝔫+=𝔱+𝔩𝔫.\mathfrak{b}_{+}^{r}\subseteq\mathfrak{l}+\mathfrak{t}\oplus\mathfrak{n}_{+}=\mathfrak{t}+\mathfrak{l}\oplus\mathfrak{n}.

Thus df+g𝔩𝔫\mathrm{d}f_{+\emptyset}^{g}\in\mathfrak{l}\oplus\mathfrak{n} and hence

Adg(df+g)Adg(𝔩𝔫)Tg.\mathrm{Ad}_{g}(\mathrm{d}f_{+\emptyset}^{g})\in\mathrm{Ad}_{g}(\mathfrak{l}\oplus\mathfrak{n})\subseteq T_{g}.

Now pick nNn\in N and n~LN\tilde{n}\in LN such that g=nw˙n~g=n\dot{w}\tilde{n}. As df+0g\mathrm{d}f_{+0}^{g} lies in 𝔱\mathfrak{t} by construction, the element Adn(df+0g)df+0g\mathrm{Ad}_{n}(\mathrm{d}f_{+0}^{g})-\mathrm{d}f_{+0}^{g} lies in 𝔫\mathfrak{n}, so that

Adnw˙n~(df+0g)Adw˙n~(df+0g)Adw˙n~(𝔫)=Adg(𝔫)Tg.\mathrm{Ad}_{n\dot{w}\tilde{n}}(\mathrm{d}f_{+0}^{g})-\mathrm{Ad}_{\dot{w}\tilde{n}}(\mathrm{d}f_{+0}^{g})\in\mathrm{Ad}_{\dot{w}\tilde{n}}(\mathfrak{n})=\mathrm{Ad}_{g}(\mathfrak{n})\subset T_{g}.

Furthermore as w(𝔱)=𝔱w(\mathfrak{t})=\mathfrak{t} normalises 𝔩𝔫\mathfrak{l}\oplus\mathfrak{n}, we have

Adw˙n~(df+0g)Adw˙(df+0g)𝔩𝔫Tg\mathrm{Ad}_{\dot{w}\tilde{n}}(\mathrm{d}f_{+0}^{g})-\mathrm{Ad}_{\dot{w}}(\mathrm{d}f_{+0}^{g})\in\mathfrak{l}\oplus\mathfrak{n}\subseteq T_{g}

so we conclude from the last three inclusions that

Adg(df+g)=Adg(df+g)+Adg(df+0g)Adw˙(df+0g)modTg.\mathrm{Ad}_{g}(\mathrm{d}f_{+}^{g})=\mathrm{Ad}_{g}(\mathrm{d}f_{+\emptyset}^{g})+\mathrm{Ad}_{g}(\mathrm{d}f_{+0}^{g})\equiv\mathrm{Ad}_{\dot{w}}(\mathrm{d}f_{+0}^{g})\mod T_{g}. (4.10)

As ff is invariant under conjugation by NN, we have f(ng)f(gn)=0f(ng)-f(gn)=0 for all nn in NN which implies that the element dfg𝔤\mathrm{d}f^{g}\in\mathfrak{g}^{*} lies in the annihilator of 𝔫\mathfrak{n}. From (4.6) one sees that dfg\mathrm{d}f_{-\emptyset}^{g} lies in

β+\𝔑𝔫β+β𝔗0𝔫β+β𝔗1𝔫β(𝔩𝔫)+𝔩+𝔩=𝔩Tg.\oplus_{\beta\in\mathfrak{R}_{+}\backslash\mathfrak{N}}\mathfrak{n}_{-\beta}+\oplus_{\beta\in\mathfrak{T}_{0}}\mathfrak{n}_{-\beta}+\oplus_{\beta\in\mathfrak{T}_{1}}\mathfrak{n}_{\beta}\subseteq(\mathfrak{l}\cap\mathfrak{n}_{-}\bigr{)}+\mathfrak{l}+\mathfrak{l}=\mathfrak{l}\subseteq T_{g}.

Combined with (4.10), the claim follows. ∎

Finally, we prove (iii):

Proof.

Recall Corollary 4.26 and Proposition 4.25; they now imply that the bracket of 𝒪G\mathcal{O}_{G_{*}} reduces to 𝒪Nw˙LNN\mathcal{O}_{N\dot{w}LN}^{N} if and only if all Hamiltonian vector fields of functions in (ι)1(𝒪Nw˙LNN)(\iota^{\sharp})^{-1}(\mathcal{O}_{N\dot{w}LN}^{N}) are tangent to Nw˙LNN\dot{w}LN. From Corollary 4.30 we deduce that Hamf(g)Tg\mathrm{Ham}_{f}(g)\in T_{g} for all f(ι)1(𝒪Nw˙LNN)f\in(\iota^{\sharp})^{-1}(\mathcal{O}_{N\dot{w}LN}^{N}) if and only if the image of

𝔱sc𝔱,tw((r0+c𝔱2)(t))(r0c𝔱2)(t)\mathfrak{t}_{\mathrm{sc}}^{*}\longrightarrow\mathfrak{t},\qquad t^{*}\longmapsto w\bigl{(}(\frac{r_{0}+c_{\mathfrak{t}}}{2})(t^{*})\bigr{)}-(\frac{r_{0}-c_{\mathfrak{t}}}{2})(t^{*})

lands inside of Tg𝔱T_{g}\cap\mathfrak{t}, which by Lemma 4.28 equals 𝔩𝔱𝔱w\mathfrak{l}\cap\mathfrak{t}\subseteq\mathfrak{t}^{w}. In other words, Hamf(g)Tg\mathrm{Ham}_{f}(g)\in T_{g} if and only if the projection of

(12(w1)(r0)+12(w+1)(c𝔱))(𝔱sc)\bigl{(}\frac{1}{2}(w-1)(r_{0})+\frac{1}{2}(w+1)(c_{\mathfrak{t}})\bigl{)}(\mathfrak{t}_{\mathrm{sc}}^{*})

to 𝔱w\mathfrak{t}_{w} is trivial in the orthogonal decomposition 𝔱=𝔱w𝔱w\mathfrak{t}=\mathfrak{t}^{w}\oplus\mathfrak{t}_{w}. Since the right-hand-side would send 𝔱scw\mathfrak{t}_{\mathrm{sc}}^{w*} to 𝔱w\mathfrak{t}^{w} and w1w-1 acts nontrivially on anything outside of 𝔱w\mathfrak{t}^{w}, it follows that r0r_{0} must map 𝔱w\mathfrak{t}^{w*} to 𝔱w\mathfrak{t}^{w}, for otherwise a nontrivial 𝔱w\mathfrak{t}_{w}-component would appear. By skew-symmetry, r0r_{0} then also preserves 𝔱w\mathfrak{t}_{w}.

By ww-invariance we may decompose c𝔱=c𝔱w+c𝔱wc_{\mathfrak{t}}=c_{\mathfrak{t}_{w}}+c_{\mathfrak{t}^{w}}. On 𝔱w\mathfrak{t}^{w} the operator w1w-1 then acts trivially, so it follows from nondegeneracy of c𝔱w=12(w+1)(c𝔱w)c_{\mathfrak{t}^{w}}=\frac{1}{2}(w+1)(c_{\mathfrak{t}^{w}}) that the image is all of 𝔱w\mathfrak{t}^{w}, so 𝔱w𝔩\mathfrak{t}^{w}\subseteq\mathfrak{l}.

Now consider r0|𝔱wr_{0}|_{\mathfrak{t}_{w}}. It needs to satisfy

(w1)(r0|𝔱w)+(w+1)(c𝔱)=0,(w-1)(r_{0}|_{\mathfrak{t}_{w}})+(w+1)(c_{\mathfrak{t}})=0,

which rewrites as

r0|𝔱w=1+w1w.r_{0}|_{\mathfrak{t}_{w}}=\frac{1+w}{1-w}.\qed

References

  • [AS19] Mohammed Abouzaid and Ivan Smith. Khovanov homology from Floer cohomology. J. Amer. Math. Soc., 32(1):1–79, 2019.
  • [BD82] A. A. Belavin and V. G. Drinfel\cprimed. Solutions of the classical Yang-Baxter equation for simple Lie algebras. Funktsional. Anal. i Prilozhen., 16(3):1–29, 96, 1982.
  • [BDG17] Mathew Bullimore, Tudor Dimofte, and Davide Gaiotto. The Coulomb branch of 3d 𝒩=4\mathcal{N}=4 theories. Comm. Math. Phys., 354(2):671–751, 2017.
  • [BFN19] Alexander Braverman, Michael Finkelberg, and Hiraku Nakajima. Coulomb branches of 3d3d 𝒩=4\mathcal{N}=4 quiver gauge theories and slices in the affine Grassmannian. Adv. Theor. Math. Phys., 23(1):75–166, 2019. With two appendices by Braverman, Finkelberg, Joel Kamnitzer, Ryosuke Kodera, Nakajima, Ben Webster and Alex Weekes.
  • [Bou68] N. Bourbaki. Éléments de mathématique. Fasc. XXXIV. Groupes et algèbres de Lie. Chapitre IV: Groupes de Coxeter et systèmes de Tits. Chapitre V: Groupes engendrés par des réflexions. Chapitre VI: systèmes de racines. Actualités Scientifiques et Industrielles, No. 1337. Hermann, Paris, 1968.
  • [Bri71] E. Brieskorn. Singular elements of semi-simple algebraic groups. In Actes du Congrès International des Mathématiciens (Nice, 1970), Tome 2, pages 279–284. 1971.
  • [BZN15] David Ben-Zvi and David Nadler. Elliptic Springer theory. Compos. Math., 151(8):1568–1584, 2015.
  • [CE48] Claude Chevalley and Samuel Eilenberg. Cohomology theory of Lie groups and Lie algebras. Trans. Amer. Math. Soc., 63:85–124, 1948.
  • [CE15] Giovanna Carnovale and Francesco Esposito. A Katsylo theorem for sheets of spherical conjugacy classes. Represent. Theory, 19:263–280, 2015.
  • [Che55] C. Chevalley. Sur certains groupes simples. Tohoku Math. J. (2), 7:14–66, 1955.
  • [Con14] Brian Conrad. Reductive group schemes. In Autour des schémas en groupes. Vol. I, volume 42/43 of Panor. Synthèses, pages 93–444. Soc. Math. France, Paris, 2014.
  • [Dav19] Dougal Davis. The elliptic Grothendieck-Springer resolution as a simultaneous log resolution of algebraic stacks. arXiv:1908.04140, August 2019.
  • [Dav20] Dougal Davis. On subregular slices of the elliptic Grothendieck-Springer resolution. arXiv:2004.13268, April 2020.
  • [dBT93] Jan de Boer and Tjark Tjin. Quantization and representation theory of finite WW algebras. Comm. Math. Phys., 158(3):485–516, 1993.
  • [DCKP92] C. De Concini, V. G. Kac, and C. Procesi. Quantum coadjoint action. J. Amer. Math. Soc., 5(1):151–189, 1992.
  • [Dix63] J. Dixmier. Représentations irréductibles des algèbres de Lie nilpotentes. An. Acad. Brasil. Ci., 35:491–519, 1963.
  • [DL76] P. Deligne and G. Lusztig. Representations of reductive groups over finite fields. Ann. of Math. (2), 103(1):103–161, 1976.
  • [Dri83] V. G. Drinfel\cprimed. Hamiltonian structures on Lie groups, Lie bialgebras and the geometric meaning of classical Yang-Baxter equations. Dokl. Akad. Nauk SSSR, 268(2):285–287, 1983.
  • [Dri89] V. G. Drinfel\cprimed. Quasi-Hopf algebras. Algebra i Analiz, 1(6):114–148, 1989.
  • [GG02] Wee Liang Gan and Victor Ginzburg. Quantization of Slodowy slices. Int. Math. Res. Not., (5):243–255, 2002.
  • [GM97] Meinolf Geck and Jean Michel. “Good” elements of finite Coxeter groups and representations of Iwahori-Hecke algebras. Proc. London Math. Soc. (3), 74(2):275–305, 1997.
  • [Gro] Ian Grojnowski. Forthcoming.
  • [GSB21] Ian Grojnowski and Nicholas Shepherd-Barron. Del Pezzo surfaces as Springer fibres for exceptional groups. Proc. Lond. Math. Soc. (3), 122(1):1–41, 2021.
  • [GW09a] Davide Gaiotto and Edward Witten. SS-duality of boundary conditions in 𝒩=4\mathcal{N}=4 super Yang-Mills theory. Adv. Theor. Math. Phys., 13(3):721–896, 2009.
  • [GW09b] Davide Gaiotto and Edward Witten. Supersymmetric boundary conditions in \scrN=4\scr N=4 super Yang-Mills theory. J. Stat. Phys., 135(5-6):789–855, 2009.
  • [GY09] K. R. Goodearl and M. Yakimov. Poisson structures on affine spaces and flag varieties. II. Trans. Amer. Math. Soc., 361(11):5753–5780, 2009.
  • [HC64] Harish-Chandra. Invariant distributions on Lie algebras. Amer. J. Math., 86:271–309, 1964.
  • [He14] Xuhua He. Geometric and homological properties of affine Deligne-Lusztig varieties. Ann. of Math. (2), 179(1):367–404, 2014.
  • [HL12] Xuhua He and George Lusztig. A generalization of Steinberg’s cross section. J. Amer. Math. Soc., 25(3):739–757, 2012.
  • [HN12] Xuhua He and Sian Nie. Minimal length elements of finite Coxeter groups. Duke Math. J., 161(15):2945–2967, 2012.
  • [Kat82] P. I. Katsylo. Sections of sheets in a reductive algebraic Lie algebra. Izv. Akad. Nauk SSSR Ser. Mat., 46(3):477–486, 670, 1982.
  • [KL88] D. Kazhdan and G. Lusztig. Fixed point varieties on affine flag manifolds. Israel J. Math., 62(2):129–168, 1988.
  • [Kos50] Jean-Louis Koszul. Homologie et cohomologie des algèbres de Lie. Bull. Soc. Math. France, 78:65–127, 1950.
  • [Kos78] Bertram Kostant. On Whittaker vectors and representation theory. Invent. Math., 48(2):101–184, 1978.
  • [KWWY14] Joel Kamnitzer, Ben Webster, Alex Weekes, and Oded Yacobi. Yangians and quantizations of slices in the affine Grassmannian. Algebra Number Theory, 8(4):857–893, 2014.
  • [Lus84] George Lusztig. Characters of reductive groups over a finite field, volume 107 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 1984.
  • [Lus15] George Lusztig. On conjugacy classes in a reductive group. In Representations of reductive groups, volume 312 of Progr. Math., pages 333–363. Birkhäuser/Springer, Cham, 2015.
  • [Mal] Wicher Malten. Forthcoming.
  • [Mal21] Wicher Malten. Minimally dominant elements of finite Coxeter groups. arXiv:2110.09266, October 2021.
  • [MR86] Jerrold E. Marsden and Tudor Ratiu. Reduction of Poisson manifolds. Lett. Math. Phys., 11(2):161–169, 1986.
  • [OR08] S. Orlik and M. Rapoport. Deligne-Lusztig varieties and period domains over finite fields. J. Algebra, 320(3):1220–1234, 2008.
  • [Pap94] Paolo Papi. A characterization of a special ordering in a root system. Proc. Amer. Math. Soc., 120(3):661–665, 1994.
  • [Pre95] Alexander Premet. Irreducible representations of Lie algebras of reductive groups and the Kac-Weisfeiler conjecture. Invent. Math., 121(1):79–117, 1995.
  • [Pre02] Alexander Premet. Special transverse slices and their enveloping algebras. Adv. Math., 170(1):1–55, 2002. With an appendix by Serge Skryabin.
  • [PT14] Alexander Premet and Lewis Topley. Derived subalgebras of centralisers and finite WW-algebras. Compos. Math., 150(9):1485–1548, 2014.
  • [Sev00] A. Sevostyanov. Quantum deformation of Whittaker modules and the Toda lattice. Duke Math. J., 105(2):211–238, 2000.
  • [Sev11] Alexey Sevostyanov. Algebraic group analogues of the Slodowy slices and deformations of Poisson WW-algebras. Int. Math. Res. Not. IMRN, (8):1880–1925, 2011.
  • [Sev19] A. Sevostyanov. Strictly transversal slices to conjugacy classes in algebraic groups. J. Algebra, 520:309–366, 2019.
  • [Sev21] Alexey Sevostyanov. Q-W-algebras, Zhelobenko operators and a proof of De Concini-Kac-Procesi conjecture. arXiv:2102.03208, February 2021.
  • [SGA3-III] Schémas en groupes. III: Structure des schémas en groupes réductifs. Séminaire de Géométrie Algébrique du Bois Marie 1962/64 (SGA 3). Dirigé par M. Demazure et A. Grothendieck. Lecture Notes in Mathematics, Vol. 153. Springer-Verlag, Berlin-New York, 1970.
  • [Slo80] Peter Slodowy. Simple singularities and simple algebraic groups, volume 815 of Lecture Notes in Mathematics. Springer, Berlin, 1980.
  • [Spr69] T. A. Springer. The unipotent variety of a semi-simple group. In Algebraic Geometry (Internat. Colloq., Tata Inst. Fund. Res., Bombay, 1968), pages 373–391. Oxford Univ. Press, London, 1969.
  • [SS06] Paul Seidel and Ivan Smith. A link invariant from the symplectic geometry of nilpotent slices. Duke Math. J., 134(3):453–514, 2006.
  • [Ste65] Robert Steinberg. Regular elements of semisimple algebraic groups. Inst. Hautes Études Sci. Publ. Math., (25):49–80, 1965.
  • [Ste68] Robert Steinberg. Lectures on Chevalley groups. Yale University, New Haven, Conn., 1968. Notes prepared by John Faulkner and Robert Wilson.
  • [STS85] Michael A. Semenov-Tian-Shansky. Dressing transformations and Poisson group actions. Publ. Res. Inst. Math. Sci., 21(6):1237–1260, 1985.
  • [STSS98] M. A. Semenov-Tian-Shansky and A. V. Sevostyanov. Drinfeld-Sokolov reduction for difference operators and deformations of \scrW{\scr W}-algebras. II. The general semisimple case. Comm. Math. Phys., 192(3):631–647, 1998.
  • [VK71] B. Ju. Veĭsfeĭler and V. G. Kac. The irreducible representations of Lie pp-algebras. Funkcional. Anal. i Priložen., 5(2):28–36, 1971.
  • [Wei83] Alan Weinstein. The local structure of Poisson manifolds. J. Differential Geom., 18(3):523–557, 1983.
  • [Wei88] Alan Weinstein. Coisotropic calculus and Poisson groupoids. J. Math. Soc. Japan, 40(4):705–727, 1988.
Mathematical Institute, University of Oxford, Oxford OX2 6GG, United Kingdom
E-mail address: [email protected]