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

\ytableausetup

centertableaux

Intersections of Deligne–Lusztig varieties and Springer fibres

Zhe Chen Department of Mathematics, Shantou University, Shantou, China [email protected]
Abstract.

In this paper we prove a direct geometric relation between Deligne–Lusztig varieties and Springer fibres in type 𝖠\mathsf{A}: For any rational unipotent element, the Springer fibre cuts out a unique component of a specific Deligne–Lusztig variety; moreover, this component forms an open dense subset of a component of the Springer fibre. This boils down to a map from the unipotent variety to the Weyl group, and combines several constructions with a combinatorial flavour (like Weyr normal forms, Robinson–Schensted correspondence, and Spaltenstein’s and Steinberg’s labellings); it also provides a geometric interpretation of a classical dimension formula of unipotent centralisers.

1. Introduction

Let 𝔾\mathbb{G} be a connected reductive group over a finite field 𝔽q\mathbb{F}_{q}, and let FF be the geometric Frobenius endomorphism on G:=𝔾×𝔽q𝔽¯qG:=\mathbb{G}\times_{\mathbb{F}_{q}}\overline{\mathbb{F}}_{q}. Fix an FF-stable Borel subgroup BGB\subseteq G and an FF-stable maximal torus TBT\subseteq B. There are two important classes of varieties lying in the flag variety G/BG/B, namely, Deligne–Lusztig varieties and Springer fibres.

Deligne–Lusztig varieties XwX_{w} (see Definition 2.1) are parametrised by wW(T):=N(T)/Tw\in W(T):=N(T)/T, and their \ell-adic cohomology (with coefficients in suitable local systems) affords all the irreducible representations of GF=𝔾(𝔽q)G^{F}=\mathbb{G}(\mathbb{F}_{q}). Meanwhile, Springer fibres u\mathcal{B}_{u} (see Definition 2.4) are parametrised by the unipotent elements uGu\in G, and their \ell-adic cohomology affords all the irreducible representations of W(T)W(T). Since their births in the seminal works [DL76] and [Spr76], respectively, these two families of varieties play crucial roles in the study of representations of Lie type groups and Weyl groups. In this paper we give a study of their relations in the case of type 𝖠\mathsf{A}.

In the remaining part of this paper we take 𝔾\mathbb{G} to be GLn\mathrm{GL}_{n}, and let BB be the standard upper Borel subgroup and TT the diagonal maximal torus.

Indeed recently we found that, for a specific unipotent uu, the intersection u,w:=Xwu\mathcal{B}_{u,w}:=X_{w}\cap\mathcal{B}_{u} appears in the study of smooth representations of the profinite group GLd(𝔽q[[π]])\mathrm{GL}_{d}(\mathbb{F}_{q}[[\pi]]), where dd is a divisor of nn (see [Che20b]); this serves the initial motivation for our attention on the relations between these two varieties. On the other hand, on the level of representations there are already striking relations found between these two constructions (see for example [Lus90] and [BV21, 1.3.5]). In this paper, instead of representations we focus on the geometric relations. We remark that a similar theme has been considered over \mathbb{C} in [Tym06], in which case Hessenberg varieties and Schubert cells took the roles of Springer fibres and Deligne–Lusztig varieties. In any case, providing the facts that both XwX_{w} and u\mathcal{B}_{u} are vital in geometric representation theory, and that they share the same ambient space G/BG/B, it is very interesting and natural to seek their geometric interactions.

We found the following surprising simple relation between the components: (See Theorem 4.1 for the formal statement.)

  • (a)

    Any Springer fibre at a rational unipotent element of GG has a component containing a component of a specific Deligne–Lusztig variety as a dense open subset; moreover, this Deligne–Lusztig component forms the whole intersection.

  • (b)

    Conversely, every component of a Deligne–Lusztig variety at an involution of a specific shape is a dense open subset of a component of some Springer fibre.

As we will see in Theorem 4.1, this component relation boils down to a map from the unipotent variety to the Weyl group.

In Section 2 we make some preparations on Deligne–Lusztig varieties and Springer fibres.

In Section 3 we give a brief recall of some ingredients used in the proof of Theorem 4.1, like Weyr normal forms, Robinson–Schensted correspondence, and Spaltenstein’s and Steinberg’s descriptions of Springer components.

In Section 4 we present the proof of Theorem 4.1, which is a careful combination of the above constructions. We also derive a geometric proof of a classical dimension formula of unipotent centralisers (see Corollary 4.2).

In Section 5 we give a few remarks, including three illustrating examples (one concerning the boundary of the theorem, one concerning a uniqueness property, and one concerning an opposite phenomenon), and a short discussion on the representations associated with u,w\mathcal{B}_{u,w}.

Throughout this paper: We use the convention notation gh=gh1=h1ghg^{h}={{}^{h^{-1}}g}=h^{-1}gh for elements g,hg,h in an algebraic group; all varieties are assumed to be reduced; by a component we always mean an irreducible component.

Acknowledgement. The author thanks George Lusztig and Alexander Stasinski for helpful comments and suggestions, and thanks Guangyi Yue for a helpful communication. During the preparation of this work the author is partially supported by the NSFC funding no.12001351.

2. Deligne–Lusztig varieties and Springer fibres

In this section we recall some basics of Deligne–Lusztig varieties and Springer fibres. The details can be found in [DL76], [Sho88], [Ste88], [Car93].

Definition 2.1.

Let wW(T)Snw\in W(T)\cong S_{n}. Then the Deligne–Lusztig variety at ww is

Xw:=L1(BwB)/B,X_{w}:=L^{-1}(BwB)/B,

where L:GGL\colon G\rightarrow G is the Lang isogeny given by gg1F(g)g\mapsto g^{-1}F(g).

Viewing G/BG/B as the variety of complete flags, one can describe XwX_{w} in the following way.

Definition 2.2.

Let

:V0V1Vnand:V0V1Vn\mathcal{F}\colon V_{0}\subseteq V_{1}\subseteq...\subseteq V_{n}\quad\textrm{and}\quad\mathcal{F}^{\prime}\colon V^{\prime}_{0}\subseteq V^{\prime}_{1}\subseteq...\subseteq V^{\prime}_{n}

be two complete flags of V:=(𝔽¯q)nV:=(\overline{\mathbb{F}}_{q})^{n}, with dimVi=dimVi=i\dim V_{i}=\dim V^{\prime}_{i}=i. We say that \mathcal{F} and \mathcal{F}^{\prime} are in relative position wW(T)=Snw\in W(T)=S_{n}, if

dimViVj=#({1,,i}{w(1),,w(j)})\dim V_{i}\cap V^{\prime}_{j}=\#\left(\{1,...,i\}\cap\{w(1),...,w(j)\}\right)

for any i,j{1,,n}i,j\in\{1,...,n\}.

  • (*)

    Let G=GLnG=\mathrm{GL}_{n} be viewed as the automorphism group of VV. Then XwX_{w} is the variety consisting of the complete flags :V0V1Vn\mathcal{F}\colon V_{0}\subseteq V_{1}\subseteq...\subseteq V_{n} such that \mathcal{F} and FF\mathcal{F} are in relative position ww, namely

    dimViFVj=#({1,,i}{w(1),,w(j)})\dim V_{i}\cap FV_{j}=\#\left(\{1,...,i\}\cap\{w(1),...,w(j)\}\right)

    for any i,j{1,,n}i,j\in\{1,...,n\}.

Proposition 2.3 (Deligne–Lusztig).

The variety XwX_{w} is a smooth locally closed subvariety of G/BG/B of pure dimension l(w)l(w), where l(w)l(w) denotes the length of ww.

Proof.

See [DL76, Page 107]. ∎

Definition 2.4.

Let uGu\in G be a unipotent element. Then the Springer fibre at uu is

u:={gBG/BugB}.\mathcal{B}_{u}:=\{gB\in G/B\mid u^{g}\in B\}.

Using the term of flags, uG/B\mathcal{B}_{u}\subseteq G/B can be viewed as the closed subvariety consisting of complete flags :V0V1Vn\mathcal{F}\colon V_{0}\subseteq V_{1}\subseteq...\subseteq V_{n} stabilised by uu (namely, uVi=ViuV_{i}=V_{i} for all ii).

Unlike XwX_{w}, usually u\mathcal{B}_{u} is singular, but one still has:

Proposition 2.5 (Spaltenstein, Steinberg).

The variety u\mathcal{B}_{u} is of pure dimension vG12dimC(u)v_{G}-\frac{1}{2}\dim C(u), where vGv_{G} denotes the number of positive roots and C(u)C(u) denotes the conjugacy class of uu.

Proof.

See e.g. [Sho88, 1.2]. ∎

In this paper we will very often take the viewpoint that elements in XwX_{w} and u\mathcal{B}_{u} are flags. We put u,w:=uXw=u×G/BXw\mathcal{B}_{u,w}:=\mathcal{B}_{u}\cap X_{w}=\mathcal{B}_{u}\times_{G/B}X_{w}.

3. Some preliminaries

In this section we recall some ingredients needed in the proof of Theorem 4.1. We first fix the notation that will be used throughout this paper:

Notation 3.1.

Let uGu\in G be a unipotent element. Then

  • J(u)=diag{J1,,Jd}J(u)=\mathrm{diag}\{J_{1},...,J_{d}\} is the standard Jordan normal form of uu, where the JiJ_{i}’s are the Jordan blocks (with non-increasing sizes).

  • rir_{i} is the size of the Jordan block JiJ_{i} (that is, JiJ_{i} is an ri×rir_{i}\times r_{i}-matrix); in particular riri+1r_{i}\geq r_{i+1}.

  • λ(u)\lambda(u) is the Young diagram associated with J(u)J(u), that is, a Young diagram whose ii-th row has rir_{i} boxes.

  • cic_{i} is the number of boxes in the ii-th column of λ(u)\lambda(u). For convenience, we also put c0=0c_{0}=0. Note that there are totally r1r_{1} columns and c1=dc_{1}=d rows.

One ingredient we would need is the so-called Weyr normal form W(u)W(u), which is a “dual” of the Jordan normal form J(u)J(u):

Definition 3.2.

For a unipotent uGLn(𝔽¯q)u\in\mathrm{GL}_{n}(\overline{\mathbb{F}}_{q}), the matrix W(u)W(u) is blocked upper triangular, and is characterised by the following rules:

  • (i)

    The ii-th diagonal block is the ci×cic_{i}\times c_{i} identity matrix.

  • (ii)

    The block just right to the ii-th diagonal block is of the form (I0)\left(\begin{smallmatrix}I\\ 0\end{smallmatrix}\right), where II denotes the ci+1×ci+1c_{i+1}\times c_{i+1}-identity matrix, and 0 denotes the zero matrix of a suitable size.

  • (iii)

    All other blocks are zero.

So W(u)W(u) is a blocked matrix of the shape

[I(I0)0000I(I0)0000I(I0)0000I(I0)0000I],\begin{bmatrix}I&\left(\begin{smallmatrix}I\\ 0\end{smallmatrix}\right)&0&0&...&0\\ 0&I&\left(\begin{smallmatrix}I\\ 0\end{smallmatrix}\right)&0&...&0\\ 0&0&I&\left(\begin{smallmatrix}I\\ 0\end{smallmatrix}\right)&...&0\\ ...&...&...&...&...&...\\ 0&...&0&0&I&\left(\begin{smallmatrix}I\\ 0\end{smallmatrix}\right)\\ 0&0&...&0&0&I\\ \end{bmatrix},

where the II’s denote some identity matrices of possibly different sizes and the 0’s denote some zero matrices of possibly different sizes.

Although Weyr normal forms and Jordan normal forms were both discovered in the second half of the 19th century, the Weyr form appears to be much lesser known; a comprehensive reference on these normal forms is [OCV11]. Note that recently there has been a (very different) application of Weyr normal forms in the representation theory of Lie type groups over local rings; see [Sta21].

Remark 3.3.

One of the main features we need from Weyr form (instead of Jordan form) is that the block sizes are with respect to the columns of Young diagrams, which allows one to combine the other elements in the proof of Theorem 4.1 in a natural way.

Proposition 3.4.

The unipotent elements uu and W(u)W(u) are in the same conjugacy class.

Proof.

See e.g. [OCV11, 2.2.2]. ∎

Another ingredient we need is the Robinson–Schensted correspondence. Recall that a very basic property of finite group representation theory is the identity ρ(dimρ)2=#H\sum_{\rho}(\dim\rho)^{2}=\#H, where HH is a finite group and ρ\rho suns over the irreducible representations. If we take H=SnH=S_{n}, then this identity can be re-written as

λ(#T(λ))2=#Sn,\sum_{\lambda}(\#T(\lambda))^{2}=\#S_{n},

where λ\lambda runs over the Young diagrams of nn boxes, and T(λ)T(\lambda) denotes the set of standard λ\lambda-tableaux (we use the convention that the numbers in a standard tableaux go increasingly from left to right and from up to down). The Robinson–Schensted correspondence, which was later generalised by Knuth to a more general situation, gives a combinatorial explanation of this identity.

Proposition 3.5 (Robinson–Schensted correspondence).

There is an explicit computable bijection between the sets

w(,):λT(λ)×T(λ)Sn,w(-,-)\colon\bigsqcup_{\lambda}T(\lambda)\times T(\lambda)\longrightarrow S_{n},

satisfying the property w(P,Q)=w(Q,P)1w(P,Q)=w(Q,P)^{-1}, given by the following algorithm:

  • (i)

    Take (P,Q)T(λ)×T(λ)(P,Q)\in T(\lambda)\times T(\lambda). If nn is in the (i,j)(i,j)-th box of QQ, then we remove this box from QQ, and denote the new tableau by QQ^{\prime}.

  • (ii)

    Suppose the number in the (i,j)(i,j)-th box of PP is nn^{\prime}. We remove this box from PP and move nn^{\prime} up by one row to replace the largest number smaller than nn^{\prime}.

  • (iii)

    Suppose the number replaced by nn^{\prime} is n′′n^{\prime\prime}, then we move n′′n^{\prime\prime} up by one row to replace the largest number smaller than n′′n^{\prime\prime}, and so on, until we replaced a number in the first row.

  • (iv)

    Denote the number been replaced from the first row by w(n)w(n), and denote the resulting tableau by PP^{\prime}.

  • (v)

    Repeat the above process for n1n-1 with the tableaux pair (P,Q)(P^{\prime},Q^{\prime}), and so on, until we find all w(n),w(n1),,w(1)w(n),w(n-1),...,w(1). Then

    w(P,Q):=(12nw(1)w(2)w(n)).w(P,Q):=\begin{pmatrix}1&2&\cdots&n\\ w(1)&w(2)&\cdots&w(n)\end{pmatrix}.

    One often use the word notation w(P,Q)=w(1)w(n)w(P,Q)=w(1)...w(n).

Proof.

See e.g. [Knu98, 5.1.4]. ∎

Besides the above two ingredients, we also need the Young tableaux labelling of components of the Springer fibre u\mathcal{B}_{u} given in [Spa76] and [Ste76]. We follow Steinberg’s description in [Ste88].

Proposition 3.6 (Tableaux labelling of components).

There is a bijection

T(λ(u)){components ofu}.T(\lambda(u))\longleftrightarrow\left\{\textrm{components of}\ \mathcal{B}_{u}\right\}.

More explicitly, for a given tableau PT(λ(u))P\in T(\lambda(u)), the corresponding component is characterised as the closure of an open subset C(P)C(P), where C(P)C(P) consists of the flags

:V0V1Vn=V\mathcal{F}\colon V_{0}\subset V_{1}\subset...\subset V_{n}=V

constructed via the following steps:

  • (I)

    Let N:=uI𝔤𝔩nN:=u-I\in\mathfrak{gl}_{n}, the nilpotent element associated with uu;

  • (II)

    if nn is in the position (i,j)=(i,ri)=(cj,j)(i,j)=(i,r_{i})=(c_{j},j) of PP, then Vn1V_{n-1} is any hyperplane satisfying that

    {NVn+KerNj1Vn1NVn+KerNjVn1;\begin{cases}NV_{n}+\mathrm{Ker}N^{j-1}\subseteq V_{n-1}\\ NV_{n}+\mathrm{Ker}N^{j}\nsubseteq V_{n-1}\\ \end{cases};
  • (III)

    once such a Vn1V_{n-1} is chosen, we repeat the above process to construct a Vn2V_{n-2} (by replacing Vn,NV_{n},N... by Vn1,N|Vn1V_{n-1},N|_{V_{n-1}}), and so on.

Proof.

See [Ste88, Section 2]. ∎

Moreover, they proved the following property of generic relative position for the components of u\mathcal{B}_{u}:

Proposition 3.7 (Generic relative position).

Let PP and QQ be two standard λ(u)\lambda(u)-tableaux. Then there is an open dense subsvariety XC(P)¯×C(Q)¯X\subseteq\overline{C(P)}\times\overline{C(Q)} such that any closed point (1,2)X(\mathcal{F}_{1},\mathcal{F}_{2})\in X has the relative position w(P,Q)w(P,Q).

Proof.

See [Ste88, Section 3] or [Spa82, II.9]. ∎

4. Distribution of components

Our main theorem is:

Theorem 4.1.

Let 𝒰\mathcal{U} be the variety of unipotent elements of GG. Then we have:

  • (a)

    There is a canonical map

    β:𝒰FW(T)\beta\colon\mathcal{U}^{F}\longrightarrow W(T)

    such that, for u𝒰Fu\in\mathcal{U}^{F}, the Springer fibre u\mathcal{B}_{u} intersects exactly one component of the Deligne–Lusztig variety Xβ(u)X_{\beta(u)}, and this component is an open dense subset of an irreducible component of u\mathcal{B}_{u}.

  • (b)

    Conversely, each component of XwX_{w}, where wW(T)=Snw\in W(T)=S_{n} is an involution of the shape w=[12][][n]w=[12...][...]...[...n] (the block [][...] means the reversion along the word), is a dense open subset of a component of some Springer fibre.

Proof.

Proof of (a).

We first prove (a) for the Weyr form W(u)W(u) of uu, in a constructible manner.

Let us construct a special λ(u)\lambda(u)-tableau 𝒯\mathcal{T} which will be critical for us: This is done by filling {1,,n}\{1,...,n\} into λ(u)\lambda(u) in the way that, first fill the 1st column of λ(u)\lambda(u) from up to down, and then the 2nd column of λ(u)\lambda(u) from up to down, and so on. So we have

(1) the i-th column of𝒯={(j<icj)+1(j<icj)+2(j<icj)+ci.{\textrm{the $i$-th column of}\ \mathcal{T}}=\begin{cases}\left(\sum_{j<i}c_{j}\right)+1\\ \left(\sum_{j<i}c_{j}\right)+2\\ ...\\ ...\\ \left(\sum_{j<i}c_{j}\right)+c_{i}\end{cases}.

(Recall that we have made the convention c0:=0c_{0}:=0.)

By the Robinson–Schensted correspondence (Proposition 3.5) this gives an involution

(2) β(u):=w(𝒯,𝒯)=i=1r1Ri,\beta(u):=w(\mathcal{T},\mathcal{T})=\prod_{i=1}^{r_{1}}R_{i},

where RiR_{i} is the reversion along the ii-th column of 𝒯\mathcal{T}:

Ri=((j<icj)+1(j<icj)+2(j<icj)+ci(j<icj)+ci(j<icj)+ci1(j<icj)+1).R_{i}=\begin{pmatrix}\left(\sum_{j<i}c_{j}\right)+1&\left(\sum_{j<i}c_{j}\right)+2&\cdots&\left(\sum_{j<i}c_{j}\right)+c_{i}\\ \left(\sum_{j<i}c_{j}\right)+c_{i}&\left(\sum_{j<i}c_{j}\right)+c_{i}-1&\cdots&\left(\sum_{j<i}c_{j}\right)+1\end{pmatrix}.

Now consider the corresponding Deligne–Lusztig variety Xβ(u)=Xw(𝒯,𝒯)X_{\beta(u)}=X_{w(\mathcal{T},\mathcal{T})}; we want to compute it using flags via (*). Let us fix a set {e1,,en}\{e_{1},...,e_{n}\} as a standard basis of VV (over 𝔽q\mathbb{F}_{q}), and view BB (resp. TT) as the standard upper triangular subgroup (resp. diagonal maximal torus) with respect to this basis. Then the Weyl group W(T)W(T) is identified as the symmetric group SnS_{n} permuting the subscripts {1,,n}\{1,...,n\} of the basis and the simple reflections are identified as the transpositions (i,i+1)(i,i+1). Then by (*), a complete flag

:V0Vn=VG/B\mathcal{F}\colon V_{0}\subseteq...\subseteq V_{n}=V\in G/B

lies in Xw(𝒯,𝒯)X_{w(\mathcal{T},\mathcal{T})} if and only if the following two conditions hold:

  • (i)

    For each cic_{i}, one has

    FVc0+c1++ci=Vc0+c1++ciFV_{c_{0}+c_{1}+...+c_{i}}=V_{c_{0}+c_{1}+...+c_{i}}

    (i.e. the spaces Vc0++ciV_{c_{0}+...+c_{i}} are FF-stable);

  • (ii)

    for any s,ts,t in the interval (c0++ci1,c0++ci)(c_{0}+...+c_{i-1},c_{0}+...+c_{i}), one has

    dimVsFVt=#({1,,s}{w(𝒯,𝒯)(1),,w(𝒯,𝒯)(t)}).\dim V_{s}\cap FV_{t}=\#\left(\{1,...,s\}\cap\left\{w(\mathcal{T},\mathcal{T})(1),...,w(\mathcal{T},\mathcal{T})(t)\right\}\right).

For each RiR_{i}, let R¯i\overline{R}_{i} be the corresponding reversion “modulo j<icj\sum_{j<i}c_{j}”, namely

R¯i=(12cicici11).\overline{R}_{i}=\begin{pmatrix}1&2&\cdots&c_{i}\\ c_{i}&c_{i}-1&\cdots&1\end{pmatrix}.

Then the condition (ii) can be re-written as:

  • (ii)’

    For any s,ts,t in the interval (c0++ci1,c0++ci)(c_{0}+...+c_{i-1},c_{0}+...+c_{i}), one has

    dimV¯sFV¯t=#({1,2,,sj<icj}{R¯i(1),R¯i(2),,R¯i(tj<icj)}),\dim\overline{V}_{s}\cap F\overline{V}_{t}=\#\left(\left\{1,2,...,s-\sum_{j<i}c_{j}\right\}\bigcap\left\{\overline{R}_{i}(1),\overline{R}_{i}(2),...,\overline{R}_{i}\left(t-\sum_{j<i}c_{j}\right)\right\}\right),

    where V¯s:=Vs/Vc0++ci1\overline{V}_{s}:=V_{s}/V_{c_{0}+...+c_{i-1}} and V¯t:=Vt/Vc0++ci1\overline{V}_{t}:=V_{t}/V_{c_{0}+...+c_{i-1}}.

Note that, considering in the quotient space Vc0++ci/Vc0++ci1V_{c_{0}+...+c_{i}}/V_{c_{0}+...+c_{i-1}}, the condition (ii)’ can be viewed as the flag condition for the Deligne–Lusztig variety XR¯iX_{\overline{R}_{i}} of

G(i):=GL(Vc0++ci/Vc0++ci1)G(i):=\mathrm{GL}(V_{c_{0}+...+c_{i}}/V_{c_{0}+...+c_{i-1}})

at R¯i\overline{R}_{i}. Therefore, the conditions (i) and (ii)’ tell us that, to construct a flag :V0Vn\mathcal{F}\colon V_{0}\subseteq...\subseteq V_{n} in Xw(𝒯,𝒯)X_{w(\mathcal{T},\mathcal{T})} is the same as to do the following two steps:

  • (iii)

    Construct an FF-stable partial flag

    {0}=Vc0Vc0+c1Vc0+c1+c2Vc0++cr1=V\{0\}=V_{c_{0}}\subseteq V_{c_{0}+c_{1}}\subseteq V_{c_{0}+c_{1}+c_{2}}\subseteq...\subseteq V_{c_{0}+...+c_{r_{1}}}=V

    of spaces of dimensions c0,c0+c1,,c0++cr1=nc_{0},c_{0}+c_{1},...,c_{0}+...+c_{r_{1}}=n, which is the same as to choose a point in (G/P)F=GF/PF(G/P)^{F}=G^{F}/P^{F}, where PP is the standard parabolic subgroup fixing the partial flag

    e1,,ec1e1,,ec1+c2e1,,ec1++cr1=en=V;\langle e_{1},...,e_{c_{1}}\rangle\subseteq\langle e_{1},...,e_{c_{1}+c_{2}}\rangle\subseteq...\subseteq\langle e_{1},...,e_{c_{1}+...+c_{r_{1}}}=e_{n}\rangle=V;
  • (iv)

    for each quotient space Vc0++ci/Vc0++ci1V_{c_{0}+...+c_{i}}/V_{c_{0}+...+c_{i-1}}, take a point in the Deligne–Lusztig variety XR¯iX_{\overline{R}_{i}}.

Therefore we get a decomposition

(3) Xw(𝒯,𝒯)(G/P)FXR¯1××XR¯r1X_{w(\mathcal{T},\mathcal{T})}\cong\bigsqcup_{(G/P)^{F}}X_{\overline{R}_{1}}\times...\times X_{\overline{R}_{r_{1}}}

into closed subvarieties. (To fix a partial part of a complete flag in the form given in the step (iii) is the same as to require a Borel subgroup to be inside a fixed parabolic subgroup, hence constitute a closed condition.) Note that R¯i\overline{R}_{i} corresponds to the longest element of the Weyl group of G(i)G(i), so XR¯iX_{\overline{R}_{i}} is an open subvariety of the flag variety of G(i)G(i), hence irreducible. Thus the disjoint union (3) is actually the component decomposition of Xw(𝒯,𝒯)X_{w(\mathcal{T},\mathcal{T})}, and GFG^{F} acts transitively on the components by translating the partial flag corresponding to PP.

Denote by C1C_{1} the component of Xw(𝒯,𝒯)X_{w(\mathcal{T},\mathcal{T})} corresponding to 1PFGF/PF1\cdot P^{F}\in G^{F}/P^{F}. By construction C1C_{1} consists of the flags :V0Vn\mathcal{F}\colon V_{0}\subseteq...\subseteq V_{n} satisfying that

  • (v)

    Vc0++ci=e1,,ec0++ciV_{c_{0}+...+c_{i}}=\langle e_{1},...,e_{c_{0}+...+c_{i}}\rangle for all ii;

  • (vi)

    ==(ii).

We shall show that C1C(𝒯)C_{1}\subseteq C(\mathcal{T}), where the latter is an open subset of the component of W(u)\mathcal{B}_{W(u)} corresponding to 𝒯\mathcal{T} (see Proposition 3.6). Consider the nilpotent element N=W(u)IN=W(u)-I, which is a blocked matrix of shape

[0(I0)00000(I0)00000(I0)00000(I0)00000],\begin{bmatrix}0&\left(\begin{smallmatrix}I\\ 0\end{smallmatrix}\right)&0&0&...&0\\ 0&0&\left(\begin{smallmatrix}I\\ 0\end{smallmatrix}\right)&0&...&0\\ 0&0&0&\left(\begin{smallmatrix}I\\ 0\end{smallmatrix}\right)&...&0\\ ...&...&...&...&...&...\\ 0&...&0&0&0&\left(\begin{smallmatrix}I\\ 0\end{smallmatrix}\right)\\ 0&0&...&0&0&0\\ \end{bmatrix},

where the II’s are the identity matrices of sizes c2,c3,,cr1c_{2},c_{3},...,c_{r_{1}} and the 0’s are the zero matrices of possibly different suitable sizes. Note that

NVc0++ciVc0++ci1NV_{c_{0}+...+c_{i}}\subseteq V_{c_{0}+...+c_{i-1}}

for any flag :V0Vn\mathcal{F}\colon V_{0}\subseteq...\subseteq V_{n} in C1C_{1}. Thus by Proposition 3.6, to show that C1C(𝒯)C_{1}\subseteq C(\mathcal{T}), it is sufficient to show that every C1\mathcal{F}\in C_{1} satisfies:

  • (vii)

    If ss is in the interval (c0++ci1,c0++ci)(c_{0}+...+c_{i-1},c_{0}+...+c_{i}), then

    {KerNi1|Vs+1VsKerNi|Vs+1Vs;\begin{cases}\mathrm{Ker}N^{i-1}|_{V_{s+1}}\subseteq V_{s}\\ \mathrm{Ker}N^{i}|_{V_{s+1}}\nsubseteq V_{s}\\ \end{cases};
  • (viii)

    if s=c0++cis=c_{0}+...+c_{i} (with i<r1i<r_{1}), then

    {KerNi|Vs+1VsKerNi+1|Vs+1Vs.\begin{cases}\mathrm{Ker}N^{i}|_{V_{s+1}}\subseteq V_{s}\\ \mathrm{Ker}N^{i+1}|_{V_{s+1}}\nsubseteq V_{s}\\ \end{cases}.

Actually this is clear, because a direct blocked-matrix computation gives that

KerNi=Vc0++ci\mathrm{Ker}N^{i}=V_{c_{0}+...+c_{i}}

for i{1,,r1}\forall i\in\{1,...,r_{1}\}. So C1C(𝒯)C_{1}\subseteq C(\mathcal{T}).

Next we show that W(u)\mathcal{B}_{W(u)} does not intersect any other component of Xw(𝒯,𝒯)X_{w(\mathcal{T},\mathcal{T})}, namely,

C1=W(u),w(𝒯,𝒯).C_{1}=\mathcal{B}_{W(u),w(\mathcal{T},\mathcal{T})}.

By the decomposition (3), it is sufficient to show that, if a flag :V0Vn=V\mathcal{F}\colon V_{0}\subseteq...\subseteq V_{n}=V is stabilised by uu (or equivalently, NVi+1ViNV_{i+1}\subseteq V_{i}, i\forall i) and satisfies the condition (ii), then it automatically satisfies the condition (v), i.e.  Vc0++ci=e1,,ec0++ciV_{c_{0}+...+c_{i}}=\langle e_{1},...,e_{c_{0}+...+c_{i}}\rangle for all ii. We first prove this for Vc1V_{c_{1}}. Note that, for any z{1,,c11}z\in\{1,...,c_{1}-1\}, the condition (ii) implies that there is an internal direct sum decomposition

(4) Vc1=VzFVc1z,V_{c_{1}}=V_{z}\oplus FV_{c_{1}-z},

which gives an internal direct sum decomposition of Vz+1V_{z+1}

Vz+1=VzFv,V_{z+1}=V_{z}\oplus F\langle v^{\prime}\rangle,

where vv^{\prime} is some non-zero vector in Vc1zV_{c_{1}-z}. So by NVz+1VzNV_{z+1}\subseteq V_{z} we get

(5) NVz+1=NVz+NFvVz.NV_{z+1}=NV_{z}+NF\langle v^{\prime}\rangle\subseteq V_{z}.

Since NN is FF-stable, (5) implies that

FNv=NFvVz;FNv^{\prime}=NFv^{\prime}\in V_{z};

however, by the internal decomposition (4), this happens if and only if Nv=0Nv^{\prime}=0. Thus from (5) we see that NVz+1=NVzNV_{z+1}=NV_{z} for any z{1,,c11}z\in\{1,...,c_{1}-1\}. Therefore

NVc1=NVc11==NV1=0,NV_{c_{1}}=NV_{c_{1}-1}=...=NV_{1}=0,

which gives that Vc1=e1,e2,,ec1V_{c_{1}}=\langle e_{1},e_{2},...,e_{c_{1}}\rangle. Now, suppose Vc0++ci=e1,,ec0++ciV_{c_{0}+...+c_{i}}=\langle e_{1},...,e_{c_{0}+...+c_{i}}\rangle for some ii, then by applying the above argument to the quotient space Vc0++ci+1/Vc0++ciV_{c_{0}+...+c_{i+1}}/V_{c_{0}+...+c_{i}} we also get Vc0++ci+1=e1,,ec0++ci+1V_{c_{0}+...+c_{i+1}}=\langle e_{1},...,e_{c_{0}+...+c_{i+1}}\rangle. So by induction the condition (v) holds, and we conclude that W(u)\mathcal{B}_{W(u)} intersects Xw(𝒯,𝒯)X_{w(\mathcal{T},\mathcal{T})} at exactly the component C1C_{1}, or more precisely,

(6) C1=W(u),w(𝒯,𝒯)=C(𝒯)Xw(𝒯,𝒯).C_{1}=\mathcal{B}_{W(u),w(\mathcal{T},\mathcal{T})}=C(\mathcal{T})\cap X_{w(\mathcal{T},\mathcal{T})}.

It remains to discuss the openness of C1C_{1} in the component C(𝒯)¯\overline{C(\mathcal{T})}. First note that, since our NN is FF-stable, the step (II) in Proposition 3.6 implies that

FC(𝒯)=C(𝒯).F{C(\mathcal{T})}={C(\mathcal{T})}.

So there is a graph embedding

(Id,F):C(𝒯)C(𝒯)×C(𝒯)G/B×G/B.(\mathrm{Id},F)\colon{C(\mathcal{T})}\longrightarrow{C(\mathcal{T})}\times{C(\mathcal{T})}\subseteq G/B\times G/B.

It is well-known that (see e.g. [Car93, 7.7]) the subsets DvG/B×G/BD_{v}\subseteq G/B\times G/B of pairs at various relative positions vv give a finite stratification of G/B×G/BG/B\times G/B into locally closed subvarieties, so their intersections with C(𝒯)×C(𝒯){C(\mathcal{T})}\times{C(\mathcal{T})} also give such a stratification of C(𝒯)×C(𝒯){C(\mathcal{T})}\times{C(\mathcal{T})}; let us denote the unique dense open strata of C(𝒯)×C(𝒯){C(\mathcal{T})}\times{C(\mathcal{T})} by XX. Then Proposition 3.7 implies that every pair in XX is in the relative position w(𝒯,𝒯)w(\mathcal{T},\mathcal{T}). In particular we see that

(Id,F)1(X)=C(𝒯)Xw(𝒯,𝒯),(\mathrm{Id},F)^{-1}(X)={C(\mathcal{T})}\cap X_{w(\mathcal{T},\mathcal{T})},

and that this is an open subvariety of C(𝒯)C(\mathcal{T}) (and hence of C(𝒯)¯\overline{C(\mathcal{T})}). So, as Xw(𝒯,𝒯)X_{w(\mathcal{T},\mathcal{T})} is pure dimensional (see Proposition 2.3), by (6) we conclude that C1C_{1} is a dense open subvariety of C(𝒯)¯\overline{C(\mathcal{T})} as desired.

The above proves (a) for W(u)W(u). For a general unipotent uGF=GLn(𝔽q)u\in G^{F}=\mathrm{GL}_{n}(\mathbb{F}_{q}), by Proposition 3.4 we know that

ug=W(u),u^{g}=W(u),

for some gGg\in G. Meanwhile, as both uu and W(u)W(u) are in GFG^{F}, by the fact that two matrices similar over a field extension are similar over the original field, we can take gGFg\in G^{F}. (Another way to see this is to apply the Lang–Steinberg theorem, by noting that centralisers in GLn/𝔽¯q\mathrm{GL}_{n/\overline{\mathbb{F}}_{q}} are always connected.) Thus gC1gC_{1} is still a component of Xw(𝒯,𝒯)X_{w(\mathcal{T},\mathcal{T})} and is also an open dense subset of the component

gC(𝒯)¯Wg(u)=u.g\overline{C(\mathcal{T})}\subseteq\mathcal{B}_{{{}^{g}W(u)}}=\mathcal{B}_{u}.

The uniqueness of the component is also clear by taking a conjugation. This completes the proof of (a).

Proof of (b).

For such an involution wW(T)w\in W(T), by Proposition 3.5 we see that w(𝒯,𝒯)=ww(\mathcal{T},\mathcal{T})=w for some tableau 𝒯\mathcal{T} of the form specified in (1). Let uu be any unipotent element making 𝒯\mathcal{T} a λ(u)\lambda(u)-tableau. Then, as we did in the proof of (a), there is a component C1C_{1} of XwX_{w} lying as an open dense subvariety in the component C(𝒯)¯\overline{C(\mathcal{T})} of W(u)\mathcal{B}_{W(u)}.

Now, from the discussion of Xw(𝒯,𝒯)X_{w(\mathcal{T},\mathcal{T})} given in (a) we know that the translation action of GFG^{F} on XwX_{w} is transitive on the components; in particular, each component of XwX_{w} is of the form gC1gC_{1} for some gGFg\in G^{F}, which is then an open dense subset of some component of Wg(u)\mathcal{B}_{{{}^{g}W(u)}}. This completes the proof. ∎

Recall the dimension formula:

Corollary 4.2.

Let uu be a unipotent element, then

dimZG(u)=ici2,\dim Z_{G}(u)=\sum_{i}c_{i}^{2},

where cic_{i} is the number of boxes in the ii-th column of the Young diagram of uu.

This can be proved by a matrix manipulation together with a combinatorial consideration (see [SS70, IV.1] and [Hum95, 1.2 and 1.3]); here we derive a geometric proof from the argument of Theorem 4.1.

Proof.

By Proposition 3.4 we can assume that u=W(u)u=W(u). Then from the proof of Theorem 4.1 we see that

dimu=dimXw(𝒯,𝒯).\dim\mathcal{B}_{u}=\dim X_{w(\mathcal{T},\mathcal{T})}.

By Proposition 2.3 and Proposition 2.5 this equality can be written as

n(n1)212dimC(u)=l(w(𝒯,𝒯)),\frac{n(n-1)}{2}-\frac{1}{2}\dim C(u)=l(w(\mathcal{T},\mathcal{T})),

where C(u)C(u) denotes the conjugacy class of uu. Note that, by the component decomposition in the argument of Theorem 4.1 (see (3)), we have

l(w(𝒯,𝒯))=il(R¯i)=ici(ci1)2.l(w(\mathcal{T},\mathcal{T}))=\sum_{i}l(\overline{R}_{i})=\sum_{i}\frac{c_{i}(c_{i}-1)}{2}.

So

dimZG(u)=dimGdimC(u)=n+ici(ci1)=ici2,\dim Z_{G}(u)=\dim G-\dim C(u)=n+\sum_{i}c_{i}(c_{i}-1)=\sum_{i}c_{i}^{2},

as desired. ∎

5. Examples and remarks

In this section we give some examples and remarks related to our main theorem, and give a short discussion on the representations associated with u,w=uXw\mathcal{B}_{u,w}=\mathcal{B}_{u}\cap X_{w} for rectangular uu.

1.

One may wonder that, in Theorem 4.1, does (a) applies to an arbitrary component of u\mathcal{B}_{u}, or does (b) applies to an arbitrary involution of W(T)W(T)? As illustrated below, without weakening the assertions usually the answer is no:

Example 5.1.

Let 𝔾=GL4\mathbb{G}=\mathrm{GL}_{4} and let

u=W(u)=[1010010100100001]GL4(𝔽q).u=W(u)=\begin{bmatrix}1&0&1&0\\ 0&1&0&1\\ 0&0&1&0\\ 0&0&0&1\end{bmatrix}\in\mathrm{GL}_{4}({\mathbb{F}}_{q}).

By Proposition 2.5 and Proposition 3.6, the variety u\mathcal{B}_{u} has two irreducible components, both of dimension 22, indexed by the Young tableaux

P:={ytableau}1&324andQ:={ytableau}1&234,P:=\ytableau 1&3\\ 24\quad\textrm{and}\quad Q:=\ytableau 1&2\\ 34,

respectively. The Robinson–Schensted correspondence gives two involutions,

w(P,P)=2143=(1,2)(3,4)w\left(P,P\right)=2143=(1,2)(3,4)

and

w(Q,Q)=3412=(2,3)(1,2)(3,4)(2,3).w\left(Q,Q\right)=3412=(2,3)(1,2)(3,4)(2,3).

(The second equalities are for writing the elements as reduced products of simple reflections.) First consider the component C(P)¯\overline{C(P)}. Explicitly, by computing the flag condition (*) for the Deligne–Lusztig variety X2143X_{2143}, and by computing the step (II) of Proposition 3.6 for C(P)C(P), one can see that

X2143Gr(2,4)F×(1\1(𝔽q))2X_{2143}\cong\mathrm{Gr}(2,4)^{F}\times\left(\mathbb{P}^{1}\backslash\mathbb{P}^{1}(\mathbb{F}_{q})\right)^{2}

and

C(P)(1)2,C(P)\cong(\mathbb{P}^{1})^{2},

which suggests that C(P)C(P) contains a component of X2143X_{2143} as a dense open subset, and from the argument of Theorem 4.1 we know that this is exactly the case, and this component is given as u,2143=uX2143\mathcal{B}_{u,2143}=\mathcal{B}_{u}\cap X_{2143}.

Now consider the other component C(Q)¯\overline{C(Q)}. As mentioned in the argument of Theorem 4.1, Proposition 3.7 implies that the variety of pairs of flags at relative position w=3412w=3412 cut out an open dense subset XX of C(Q)×C(Q){C(Q)}\times{C(Q)}, and so the preimage C(Q)X3412{C(Q)}\cap X_{3412} of XX along the Frobenius graph embedding

(Id,F):C(Q)C(Q)×C(Q)(\mathrm{Id},F)\colon{C(Q)}\longrightarrow{C(Q)}\times{C(Q)}

is an open subset of C(Q)¯\overline{C(Q)}; indeed, this open subset is non-empty: By fixing a standard basis {e1,e2,e3,e4}\{e_{1},e_{2},e_{3},e_{4}\} of V=𝔽¯q4V=\overline{\mathbb{F}}_{q}^{4} (over 𝔽q\mathbb{F}_{q}), one easily checks that it contains the flag

{0}{e1+xe2}{e1+xe2,e3+xe4}{e1,e2,e3+xe4}V\{0\}\subseteq\{e_{1}+xe_{2}\}\subseteq\{e_{1}+xe_{2},e_{3}+xe_{4}\}\subseteq\{e_{1},e_{2},e_{3}+xe_{4}\}\subseteq V

for any x𝔽¯q\𝔽qx\in\overline{\mathbb{F}}_{q}\backslash\mathbb{F}_{q}. This means that some component of X3412X_{3412} cuts out an open dense subset of C(Q)¯\overline{C(Q)}. However, since

dimX3412=l((2,3)(1,2)(3,4)(2,3))=4>2=dimu,\dim X_{3412}=l((2,3)(1,2)(3,4)(2,3))=4>2=\dim\mathcal{B}_{u},

and since the Deligne–Lusztig varieties partition the flag variety, the component C(Q)¯\overline{C(Q)} cannot contain any component of a Deligne–Lusztig variety as an open dense subset. On the other hand, by Proposition 2.5 and Corollary 4.2 we see that the possible dimensions of Springer fibres for GL4\mathrm{GL}_{4} are 0,1,2,3,60,1,2,3,6, so none of the components of X3412X_{3412} can be an open subset of a Springer fibre.

2.

It is a natural desire that, the map β()\beta(-) is uniquely characterised by the property given in Theorem 4.1. To make this true one shall add further conditions:

Example 5.2.

Let G=GL3G=\mathrm{GL}_{3} and let

u=W(u)=[101010001]GL3(𝔽q).u=W(u)=\begin{bmatrix}1&0&1\\ 0&1&0\\ 0&0&1\end{bmatrix}\in\mathrm{GL}_{3}(\mathbb{F}_{q}).

Then the components of u\mathcal{B}_{u} are labelled by the two tableaux of hook shape

P:={ytableau}1&32andQ:={ytableau}1&23,P:=\ytableau 1&3\\ 2\quad\textrm{and}\quad Q:=\ytableau 1&2\\ 3,

corresponding to the simple reflections (1,2)(1,2) and (2,3)(2,3), via the Robinson–Schensted correspondence, respectively. The three varieties u\mathcal{B}_{u}, X(1,2)X_{(1,2)}, and X(2,3)X_{(2,3)}, are all of dimension 11. From the argument of Theorem 4.1 we know that C(P)C(P) contains a component of X(1,2)X_{(1,2)} as a dense open subset and u\mathcal{B}_{u} does not intersect other components of X(1,2)X_{(1,2)}. Let us consider X(2,3)X_{(2,3)}. Given a flag V0V1V2V3VV_{0}\subseteq V_{1}\subseteq V_{2}\subseteq V_{3}\subseteq V, by direct computations with the flag condition (*) and the condition that N=uIN=u-I takes ViV_{i} into Vi1V_{i-1}, one sees that u\mathcal{B}_{u} intersects X(2,3)X_{(2,3)} only at the component (of X(2,3)X_{(2,3)}) consisting of the flags

{0}{e1}{e1,e2+λe3}{e1,e2,e3}=V,\{0\}\subseteq\{e_{1}\}\subseteq\{e_{1},e_{2}+\lambda e_{3}\}\subseteq\{e_{1},e_{2},e_{3}\}=V,

where λ\lambda runs over 𝔽¯q\𝔽q\overline{\mathbb{F}}_{q}\backslash\mathbb{F}_{q}. Meanwhile, it follows from the step (II) of Proposition 3.6 that this component is contained in C(Q)C(Q). So, at this uu, there are two choices of the value of β\beta fulfilling the requirement in Theorem 4.1.

Thus we hope to state here a question:

Question 5.3.

Is there a geometric property (in addition to the one asserted in Theorem 4.1, but without referencing to the explicit construction (2) given in its argument) making the map β\beta unique?

3.

Let w0w_{0} be the longest element of W(T)W(T), then Xw0X_{w_{0}} contributes a generic (i.e. open dense) part of the flag variety. Quite opposite to the component containment relation in Theorem 4.1, Springer fibres missed this “largest” Deligne–Lusztig variety at all:

Example 5.4.

In this example we show that u,w0=uXw0\mathcal{B}_{u,w_{0}}=\mathcal{B}_{u}\cap X_{w_{0}} is always empty unless u=1u=1. Suppose that u,w0\mathcal{B}_{u,w_{0}} is non-empty; let u,w0\mathcal{F}\in\mathcal{B}_{u,w_{0}} be a point in the component C(P)¯\overline{C(P)} for some λ(u)\lambda(u)-tableau PP. Consider the Frobenius graph embedding

(Id,F):C(P)¯C(P)¯×C(P)¯.(\mathrm{Id},F)\colon\overline{C(P)}\longrightarrow\overline{C(P)}\times\overline{C(P)}.

(Here FF preserves C(P)¯\overline{C(P)} because FF preserves C(P)C(P) and FF is a homeomorphism.) By Proposition 3.7 and by considering Bruhat order we get

``the relative position of(,F)"=w0w(P,P),``\textrm{the relative position of}\ (\mathcal{F},F\mathcal{F})"=w_{0}\leq w(P,P),

which in turn implies that w(P,P)=w0w(P,P)=w_{0}. So, via the Robinson–Schensted correspondence we see that λ(u)\lambda(u) must be a single column diagram, that is, u=1u=1, in which case u\mathcal{B}_{u} is the whole flag variety.

4.

Recall that (see Notation 3.1) we have let J(u)J(u) be the Jordan normal form of uu, dd the number of Jordan blocks in J(u)J(u), rir_{i} the sizes of each Jordan block, and λ(u)\lambda(u) the associated Young diagram.

Definition 5.5.

A unipotent element uGu\in G is called rectangular, if the Young diagram λ(u)\lambda(u) is rectangular.

Note that if uu is rectangular, then dd (=ci,i=c_{i},\forall i) is a divisor of nn and n/d=rin/d=r_{i} for all ii; when this is the case we denote rir_{i} by rr.

As mentioned in the introduction, our original focus on the relations between Deligne–Lusztig varieties and Springer fibres comes from the smooth representation theory of the profinite group GLd(𝔽q[[π]])\mathrm{GL}_{d}(\mathbb{F}_{q}[[\pi]]). Indeed, if u𝒰Fu\in\mathcal{U}^{F} is rectangular, then the finite quotient group GLd(𝔽q[[π]]/πr)\mathrm{GL}_{d}(\mathbb{F}_{q}[[\pi]]/\pi^{r}) acts on u,w\mathcal{B}_{u,w}, because GLd(𝔽¯q[[π]]/πr)\mathrm{GL}_{d}(\overline{\mathbb{F}}_{q}[[\pi]]/\pi^{r}) is naturally isomorphic to the GG-centraliser of

W(u)=[IdId0000IdId0000Id00000IdId0000Id],W(u)=\begin{bmatrix}I_{d}&I_{d}&0&0&...&0\\ 0&I_{d}&I_{d}&0&...&0\\ 0&0&I_{d}&0&...&0\\ ...&...&...&...&...&...\\ 0&...&0&0&I_{d}&I_{d}\\ 0&0&...&0&0&I_{d}\\ \end{bmatrix},

where IdI_{d} denotes the d×dd\times d identity matrix, by the ring injection from Md(𝔽¯q[[π]]/πr)M_{d}(\overline{\mathbb{F}}_{q}[[\pi]]/\pi^{r}) to Mn(𝔽¯q){M}_{n}(\overline{\mathbb{F}}_{q}):

(7) A0+A1π++Ar1πr1[A0A1Ar2Ar10A0A1Ar200A0A1000A0],A_{0}+A_{1}\pi+...+A_{r-1}\pi^{r-1}\longmapsto\begin{bmatrix}A_{0}&A_{1}&...&A_{r-2}&A_{r-1}\\ 0&A_{0}&A_{1}&...&A_{r-2}\\ ...&...&...&...&...\\ 0&...&0&A_{0}&A_{1}\\ 0&0&...&0&A_{0}\end{bmatrix},

where AiMd(𝔽¯q)A_{i}\in M_{d}(\overline{\mathbb{F}}_{q}). (See [Che20a, 5.3] and [Che20b, 4.6]; note that the notation used there is different up to a blocked transpose.) Thus we get a virtual representation

Ru,w:=i(1)iHci(u,w,¯)R_{u,w}:=\sum_{i}(-1)^{i}H_{c}^{i}(\mathcal{B}_{u,w},\overline{\mathbb{Q}}_{\ell})

of GLd(𝔽q[[π]]/πr)\mathrm{GL}_{d}(\mathbb{F}_{q}[[\pi]]/\pi^{r}). Here Hci(,¯)H_{c}^{i}(-,\overline{\mathbb{Q}}_{\ell}) denotes the ii-th compactly supported \ell-adic cohomology with \ell a prime not equal to char(𝔽q)\mathrm{char}(\mathbb{F}_{q}). Note that if n=dn=d, then this construction gives the unipotent representations of GLn(𝔽q)\mathrm{GL}_{n}(\mathbb{F}_{q}) in the sense of [DL76, 7.8].

Definition 5.6.

A representation of GLd(𝔽q[[π]]/πr)\mathrm{GL}_{d}(\mathbb{F}_{q}[[\pi]]/\pi^{r}), where r2r\geq 2, is called primitive, if it does not factor through GLd(𝔽q[[π]]/πr1)\mathrm{GL}_{d}(\mathbb{F}_{q}[[\pi]]/\pi^{r-1}).

For example, when u𝒰Fu\in\mathcal{U}^{F} is rectangular with d,r2d,r\geq 2, from [Che20b, 4.8] we known that RW(u),wR_{W(u),w} is a primitive representation of GLd(𝔽q[[π]]/πr)\mathrm{GL}_{d}(\mathbb{F}_{q}[[\pi]]/\pi^{r}) if w=(1,,z)w=(1,...,z) is a cycle with z<dz<d.

Question 5.7.

For a given rectangular u𝒰Fu\in\mathcal{U}^{F} (resp. Weyl element wW(T)w\in W(T)), is there a reasonable characterisation of the Weyl element ww (resp. rectangular u𝒰Fu\in\mathcal{U}^{F}) making Ru,wR_{u,w} primitive?

We note that there is a uniform description of the representations Ru,wR_{u,w} for various rectangular u𝒰Fu\in\mathcal{U}^{F}, using a complex on GG: First consider the character-sheaf type diagram (see [Lus85])

Xw{X_{w}}Zw{Z_{w}}G,{G,}b\scriptstyle{b}a\scriptstyle{a}

where

Zw:={(g,xB)G×L1(BwB)/BgxB},Z_{w}:=\{(g,xB)\in G\times L^{-1}(BwB)/B\mid g^{x}\in B\},

and a,ba,b are the natural projections; this diagram is GFG^{F}-equivariant, where GFG^{F} acts on ZwZ_{w} by h(g,x)=(hgh1,hx)h\cdot(g,x)=(hgh^{-1},hx), on GG by left conjugation, and on XwX_{w} by left multiplication. Then the character-sheaf type complex

K:=Ra!b¯Dcb(G,¯)K:=Ra_{!}b^{*}\overline{\mathbb{Q}}_{\ell}\in D^{b}_{c}(G,\overline{\mathbb{Q}}_{\ell})

encodes all Ru,wR_{u,w}:

Proposition 5.8.

If u𝒰Fu\in\mathcal{U}^{F} is rectangular, then Ru,w=i(1)ii(K)uR_{u,w}=\sum_{i}(-1)^{i}\mathcal{H}^{i}(K)_{u}.

Proof.

Note that u,wa1(u)\mathcal{B}_{u,w}\cong a^{-1}(u), so the assertion formally follows from the proper base change along {u}G\{u\}\hookrightarrow G. ∎

We end with a (non-)smoothness property of the above diagram.

Proposition 5.9.

The variety ZwZ_{w} and the morphism bb are smooth. However, the morphism aa is smooth only for n=1n=1.

Proof.

We follow the idea of [Lus85, 2.5.2] to use a faithful flat descent. Let Zw~\widetilde{Z_{w}} be the base change of ZwG×L1(BwB)/BZ_{w}\subseteq G\times L^{-1}(BwB)/B along the morphism

G×L1(BwB)G×L1(BwB)/B.G\times L^{-1}(BwB)\longrightarrow G\times L^{-1}(BwB)/B.

This is faithfully flat since BB is solvable (hence the quotient L1(BwB)L1(BwB)/BL^{-1}(BwB)\rightarrow L^{-1}(BwB)/B admits local sections). Then by [GD67, 17.7.7] it suffices to prove that the variety Zw~={(g,x)G×L1(BwB)gxB}G×G\widetilde{Z_{w}}=\{(g,x)\in G\times L^{-1}(BwB)\mid g^{x}\in B\}\subseteq G\times G and the morphism b~:Zw~Xw\widetilde{b}\colon\widetilde{Z_{w}}\rightarrow X_{w} (extending bb) are smooth. Applying the variable change y=gxy=g^{x} we get

Zw~{(y,x)G×L1(BwB)yB}=B×L1(BwB),\widetilde{Z_{w}}\cong\{(y,x)\in G\times L^{-1}(BwB)\mid y\in B\}=B\times L^{-1}(BwB),

and b~\widetilde{b} reads as (y,x)xB(y,x)\mapsto xB. As the Lang morphism LL is étale, the assertion on ZwZ_{w} and bb follows.

For the assertion on aa, again by the faithful flat descent it suffices to show that the morphism Zw~G\widetilde{Z_{w}}\rightarrow G given by

a~:(y,x)B×L1(BwB)xyx1\widetilde{a}\colon(y,x)\in B\times L^{-1}(BwB)\longmapsto xyx^{-1}

is not smooth. Actually this morphism is not flat: Note that the fibre of a~\widetilde{a} at any central element cZ(G)c\in Z(G) is {c}×L1(BwB)B×L1(BwB)\{c\}\times L^{-1}(BwB)\subseteq B\times L^{-1}(BwB). So, if a~\widetilde{a} is flat then we have (see e.g. [Liu06, 4.3.12])

dimL1(BwB)=dimB+dimL1(BwB)dimG,\dim L^{-1}(BwB)=\dim B+\dim L^{-1}(BwB)-\dim G,

which is impossible unless n=1n=1. ∎

In the above argument, note that for any bBb\in B, the elements (y,x)(y,x) and (b1yb,xb)(b^{-1}yb,xb) have the same image under a~\widetilde{a}. So, the fibre of a~\widetilde{a} at a closed point is either empty or has dimension at least dimB\dim B, which implies that (via [Liu06, 4.3.12])

dimBa~1(xyx1)=dimB+dimL1(BwB)dimG\dim B\leq\widetilde{a}^{-1}(xyx^{-1})=\dim B+\dim L^{-1}(BwB)-\dim G

if a~\widetilde{a} is flat at (y,x)(y,x). Thus actually a~\widetilde{a} cannot be flat at any point unless ww is the longest element w0w_{0}.

References

  • [BV21] Roman Bezrukavnikov and Yakov Varshavsky. Affine Springer fibers and depth zero L-packets. arXiv preprint arXiv:2104.13123, 2021.
  • [Car93] Roger W. Carter. Finite groups of Lie type. Wiley Classics Library. John Wiley & Sons Ltd., Chichester, 1993. Conjugacy classes and complex characters, Reprint of the 1985 original, A Wiley-Interscience Publication.
  • [Che20a] Zhe Chen. Flags and orbits of connected reductive groups over local rings. Math. Ann., 376(3-4):1449–1466, 2020.
  • [Che20b] Zhe Chen. Twisting operators and centralisers of lie type groups over local rings. arXiv preprint arXiv:2006.02145, 2020.
  • [DL76] Pierre Deligne and George Lusztig. Representations of reductive groups over finite fields. Ann. of Math. (2), 103(1):103–161, 1976.
  • [GD67] Alexander Grothendieck and Jean Dieudonné. Éléments de géométrie algébrique IV Étude locale des schémas et des morphismes de schémas. Inst. Hautes Études Sci. Publ. Math., 1964-1967.
  • [Hum95] James E. Humphreys. Conjugacy classes in semisimple algebraic groups, volume 43 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 1995.
  • [Knu98] Donald E. Knuth. The art of computer programming. Vol. 3. Addison-Wesley, Reading, MA, 1998. Sorting and searching, Second edition.
  • [Liu06] Qing Liu. Algebraic geometry and arithmetic curves, volume 6 of Oxford Graduate Texts in Mathematics. Oxford University Press, 2006.
  • [Lus85] George Lusztig. Character sheaves I. Adv. in Math., 56(3):193 – 237, 1985.
  • [Lus90] George Lusztig. Green functions and character sheaves. Ann. of Math. (2), 131(2):355–408, 1990.
  • [OCV11] Kevin C. O’Meara, John Clark, and Charles I. Vinsonhaler. Advanced topics in linear algebra. Oxford University Press, Oxford, 2011. Weaving matrix problems through the Weyr form.
  • [Sho88] Toshiaki Shoji. Geometry of orbits and Springer correspondence. Astérisque, (168):9, 61–140, 1988. Orbites unipotentes et représentations, I.
  • [Spa76] Nicolas Spaltenstein. The fixed point set of a unipotent transformation on the flag manifold. Nederl. Akad. Wetensch. Proc. Ser. A, 38(5):452–456, 1976.
  • [Spa82] Nicolas Spaltenstein. Classes unipotentes et sous-groupes de Borel, volume 946 of Lecture Notes in Mathematics. Springer-Verlag, Berlin-New York, 1982.
  • [Spr76] Tonny A. Springer. Trigonometric sums, Green functions of finite groups and representations of Weyl groups. Invent. Math., 36:173–207, 1976.
  • [SS70] Tonny A. Springer and Robert Steinberg. Conjugacy classes. In Seminar on Algebraic Groups and Related Finite Groups (The Institute for Advanced Study, Princeton, N.J., 1968/69), Lecture Notes in Mathematics, Vol. 131, pages 167–266. Springer, Berlin, 1970.
  • [Sta21] Alexander Stasinski. Representations of SLn{\rm SL}_{n} over finite local rings of length two. J. Algebra, 566:119–135, 2021.
  • [Ste76] Robert Steinberg. On the desingularization of the unipotent variety. Invent. Math., 36:209–224, 1976.
  • [Ste88] Robert Steinberg. An occurrence of the Robinson-Schensted correspondence. J. Algebra, 113(2):523–528, 1988.
  • [Tym06] Julianna S. Tymoczko. Linear conditions imposed on flag varieties. Amer. J. Math., 128(6):1587–1604, 2006.