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

The centralizer of a classical group and Bruhat–Tits buildings

Daniel Skodlerack
Abstract

Let GG be a unitary group defined over a non-Archimedean local field of odd residue characteristic and let HH be the centralizer of a semisimple rational Lie algebra element of G.G. We prove that the Bruhat–Tits building 𝔅1(H)\mathfrak{B}^{1}(H) of HH can be affinely and GG-equivariantly embedded in the Bruhat–Tits building 𝔅1(G)\mathfrak{B}^{1}(G) of GG so that the Moy–Prasad filtrations are preserved. The latter property forces uniqueness in the following way. Let jj and jj^{\prime} be maps from 𝔅1(H)\mathfrak{B}^{1}(H) to 𝔅1(G)\mathfrak{B}^{1}(G) which preserve the Moy–Prasad filtrations. We prove that if there is no split torus in the center of the connected component of HH then jj and jj^{\prime} are equal, and in general if both maps are affine and satisfy a mild equivariance condition they differ up to a translation of 𝔅1(H).\mathfrak{B}^{1}(H).

1 Introduction

The subject of this article is a functoriality question for maps between Bruhat–Tits buidlings which is connected with the representation theory of classical groups. Embeddings of buildings of reductive groups have previously been studied for example by Landvogt [Lan00] and Prasad and Yu [PY02]. The aim of this article is to show to what extend the property of preserving the Moy–Prasad filtrations determines the choice of an embedding uniquely. It completes recent works of Broussous, Lemaire and Stevens [BL02], [BS09], which have applications in representation theory.

More precisely Bushnell’s and Kutzko’s strategy in their theory of types for the classification of irreducible smooth representations of GLn(k)\text{GL}_{n}(k) in [BK93] is applied to other classical groups defined over a non-Archimedean local field k.k. In [Séc04], [Séc05a], [Séc05b], [SS08], [BSS10] and [SS11] Sécherre together with Broussous and Stevens gave the classification for GLn(D),\text{GL}_{n}(D), where DD is a central finite division algebra over k.k. Further in [Ste05] Stevens constructed types for unitary groups. He applied a result of his paper with Broussous [BS09], i.e. he used an embedding of Bruhat–Tits buildings for a certain subgroup of a unitary group to apply an induction. The important property of this map is the compatibility with the Lie algebra filtrations (CLF) which by [Lem09] correspond to the Moy–Prasad filtrations [MP94]. In [BS09] the quaternion algebra case is missing and the authors proposed a uniqueness and generalization conjecture to the reader.

Let kk be a non-Archimedean local field with valuation ν\nu and of residue characteristic 2,\neq 2, and let DD be a division algebra central and finite-dimensional over kk equipped with an involution ρ,\rho, whose set of fixed points in kk is denoted by k0.k_{0}. To the classical group 𝐆:=𝐔(h),\operatorname{\mathbf{G}}:=\operatorname{\mathbf{U}}(h), i.e. the unitary group of an ϵ\epsilon-hermitian form hh on a finite-dimensional right DD-vector space V,V, is attached the Bruhat–Tits building 𝔅1(𝐆,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{G}},k_{0}) which can be described in terms of lattice functions. To every point of 𝔅1(𝐆,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{G}},k_{0}) there is attached a Lie algebra filtration 𝔤x\mathfrak{g}_{x} which is exactly the Moy–Prasad filtration. More precisely if xx is a point of 𝔅1(𝐆,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{G}},k_{0}) seen as a lattice function it can be interpreted as a point of 𝔅(𝐆𝐋D(V),k)\mathfrak{B}(\operatorname{\mathbf{GL}}_{D}(V),k) which has a Lie algebra filtration 𝔤~x\tilde{\mathfrak{g}}_{x} in EndD(V).\operatorname{End}_{D}(V). If one identifies Lie(𝐆)(k0)\operatorname{Lie}(\operatorname{\mathbf{G}})(k_{0}) with the set of skew-symmetric elements of EndD(V)\operatorname{End}_{D}(V) with respect to the adjoint involution of h,h, the filtration 𝔤x\mathfrak{g}_{x} of xx coincides with

t𝔤~x(t)Lie(𝐆)(k0).t\mapsto\tilde{\mathfrak{g}}_{x}(t)\cap\operatorname{Lie}(\operatorname{\mathbf{G}})(k_{0}).

Now we take an element βLie(𝐆)(k0)\beta\in\operatorname{Lie}(\operatorname{\mathbf{G}})(k_{0}) whose kk-algebra k[β]k[\beta] is a product of field extensions EiE_{i} of k.k. In this introduction let us assume k[β]k[\beta] to be separable over k.k.

Its centralizer in 𝐆\operatorname{\mathbf{G}} is an algebraic group 𝐇\operatorname{\mathbf{H}} defined over k0k_{0} which is a product of restrictions of scalars to k0k_{0} of classical groups 𝐇i\operatorname{\mathbf{H}}_{i} which are either general linear groups or unitary groups. In the manner of [BS09] we prove the existence of an injective affine 𝐇(k0)\operatorname{\mathbf{H}}(k_{0})-equivariant and toral map jβj_{\beta} from 𝔅1(𝐇,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{H}},k_{0}) to 𝔅1(𝐆,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{G}},k_{0}) using lattice function models. In addition jβj_{\beta} has the CLF-property, i.e.

Lie(𝐇)(k0)𝔤jβ(y)\operatorname{Lie}(\operatorname{\mathbf{H}})(k_{0})\cap\mathfrak{g}_{j_{\beta}(y)}

is the Lie algebra flitration 𝔥x\mathfrak{h}_{x} of x.x. This article solves the following problem: To what extent does the CLF-property determine jβ?j_{\beta}?

In order to give an answer to this question we consider two cases.

  1. (a)

    If only unitary groups appear among the 𝐇i,\operatorname{\mathbf{H}}_{i}, none of which is k0k_{0}-isomorphic to the isotropic orthogonal group of rank one, then jβj_{\beta} is uniquely determined by CLF.

  2. (b)

    If there are no restrictions on the 𝐇i\operatorname{\mathbf{H}}_{i} then jβj_{\beta} is unique up to a translation of 𝔅1(𝐇,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{H}},k_{0}) in being affine, Z(𝐇0(k0))\operatorname{Z}(\operatorname{\mathbf{H}}^{0}(k_{0}))-equivariant and having the CLF-property.

For the proofs we use the decomposition of k[β]k[\beta] as a product of fields to restrict to the cases where k[β]k[\beta] is a field or a product of two fields interchanged by the adjoint involution of h.h. Let us call these cases atomic cases. The map jβj_{\beta} is now induced by jE1j_{E}^{-1} constructed in [BL02]. In (a) k[β]k[\beta] is a field in the atomic case. If β\beta is non-zero we follow the strategy of [BS09]. The statement follows essentially from a uniqueness statement for jE1j_{E}^{-1} given in [BS09, 10.3]. If β\beta is zero we prove that for all unitary groups except for the isotropic orthogonal group of rank one the Moy–Prasad filtration determines the point completely. For (b) to restrict to the atomic case we need further a rigidity proposition for Euclidean buildings, stated in 10.2. We use (a) to finish the proof of (b).

The whole strategy and the proofs do not require β\beta to be separable. In that case we define the building of Z𝐆(k0)(β)\operatorname{Z}_{\operatorname{\mathbf{G}}(k_{0})}(\beta) in view of [BS09] using lattice functions and we work mainly with the rational points instead of the algebraic groups.

The article is structured in the following way. After preliminary notation in §2 and the model of the Bruhat–Tits building of 𝐆\operatorname{\mathbf{G}} over k0k_{0} in terms of lattice functions is explained in §3. In §4 we introduce the Moy–Prasad filtration for our purposes. We give the building of the centralizer in §5 and introduce the notion of CLF in §6 followed by the existence theorem in §7. The uniqueness theorems in §9, where no GLm\text{GL}_{m} is involved, and §11, for the general case, are prepared in the preceding sections. Lastly in §12 we show that the constructed map respects apartments.

I thank very much P. Broussous, Prof. S. Stevens and Prof. E.-W. Zink for useful hints and fruitful communications. I want to express my gratitude to the German Research Foundation, who supported this work within the framework of BMS and SFB 878.

2 Notation

We are given a division algebra DD of finite index dd and central over a non-Archimedean local field kk of odd residue characteristic. The valuation on kk and its unique extension to DD are denoted by ν.\nu. We assume that the image of ν|k\nu|_{k} is .\mbox{$\mathbb{Z}$}. Further let ρ\rho be an involution on D,D, i.e. a skew field isomorphism from DD to DopD^{op} of order one or two, in particular ρ\rho is an isometry. The fixed field of ρ\rho in kk is denoted by k0.k_{0}.

Remark 2.1

[Sch85, 10.2.2] The existence of ρ\rho implies d2.d\leq 2. If kk0,k\neq k_{0}, then d=1.d=1.

We fix an element ϵ{1,1}\epsilon\in\{1,-1\} and a non-degenerate ϵ\epsilon-hermitian form hh on an mm-dimensional right DD-vector space V,V, i.e. a \mathbb{Z}-bi-linear map hh on V×VV\times V with values in DD such that

h(v1λ1,v2λ2)=ϵρ(λ1)ρ(h(v2,v1))λ2h(v_{1}\lambda_{1},v_{2}\lambda_{2})=\epsilon\rho(\lambda_{1})\rho(h(v_{2},v_{1}))\lambda_{2}

for all λ1,λ2D\lambda_{1},\lambda_{2}\in D and v1,v2V.v_{1},v_{2}\in V. Further ρ|k\rho|_{k} extends to the adjoint involution σ\sigma of hh on EndD(V).\operatorname{End}_{D}(V). For a skew field DD^{\prime} with discrete valuation the symbols oDo_{D^{\prime}} and 𝔭D\mathfrak{p}_{D^{\prime}} denote the valuation ring and its maximal ideal respectively. We fix an algebraic closure k¯\bar{k} of k.k.

By a Bruhat–Tits building we alway understand the extended one [BT84a, 4.2.16.]. The set

G:=U(σ):={fEndD(V)|σ(f)f=idV}G:=\operatorname{U}(\sigma):=\{f\in\operatorname{End}_{D}(V)|\ \sigma(f)f=\operatorname{id}_{V}\}

is the set of k0k_{0}-rational points of an algebraic group 𝐔(σ)\operatorname{\mathbf{U}}(\sigma) defined over k0.k_{0}. We also write 𝐔(h)\operatorname{\mathbf{U}}(h) for 𝐔(σ).\operatorname{\mathbf{U}}(\sigma). We denote 𝐔(h)\operatorname{\mathbf{U}}(h) by 𝐆\operatorname{\mathbf{G}} and Lie(𝐆)(k0)\operatorname{Lie}(\operatorname{\mathbf{G}})(k_{0}) by 𝔤\mathfrak{g} and identify the latter with the set Skew(EndD(V),σ)\operatorname{Skew}(\operatorname{End}_{D}(V),\sigma) of skew-symmetric elements of EndD(V)\operatorname{End}_{D}(V) with respect to σ.\sigma. The Bruhat–Tits building of 𝔅1(𝐆,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{G}},k_{0}) is also denoted by 𝔅1(G).\mathfrak{B}^{1}(G). We repeat the strategy in [BS09] to describe the building as a set of self-dual lattice functions based on the description using norms, see [BT87].

3 Norms and lattice functions

The description of the Bruhat-Tits building in terms of norms and lattice functions needs some basic properties which are collected in this section. Proofs and more details for norms can mainly be found in [BT84b], [BT87]. For lattice functions we refer to [BS09] and [BL02].

Definition 3.1

[BT84b, 1.1] A DD-norm on VV is a map α:V{}\alpha:\ V\mbox{$\rightarrow$}\mbox{$\mathbb{R}$}\cup\{\infty\} such that

  • α(v)=\alpha(v)=\infty if and only if v=0,v=0,

  • α(vλ)=ν(λ)+α(v),\alpha(v\lambda)=\nu(\lambda)+\alpha(v),

  • α(v+w)inf{α(v),α(w)}\alpha(v+w)\geq\inf\{\alpha(v),\alpha(w)\},

for all v,wVv,w\in V and λD.\lambda\in D. Fix a DD-norm α\alpha on V.V. The dual α#\alpha^{\#} of α\alpha with respect to hh is defined to be the DD-norm

vVinf{ν(h(v,w))α(w)|wV,w0}.v\in V\mapsto\inf\{\nu(h(v,w))-\alpha(w)|\ w\in V,\ w\neq 0\}.

One calls α\alpha self-dual with respect to hh if it is equal to his dual. The set of DD-norms and self-dual DD-norms of VV are denoted by NormD1(V)\operatorname{Norm}^{1}_{D}(V) and Normh1(V)\operatorname{Norm}^{1}_{h}(V) respectively. Given a further right DD-vector space VV^{\prime} and a norm αNormD1(V)\alpha^{\prime}\in\operatorname{Norm}^{1}_{D}(V^{\prime}) the direct sum of α\alpha and α\alpha^{\prime} is defined by

(αα)(v+v):=inf{α(v),α(v)},vV,vV.(\alpha\oplus\alpha^{\prime})(v+v^{\prime}):=\inf\{\alpha(v),\alpha^{\prime}(v^{\prime})\},\ v\in V,\ v^{\prime}\in V^{\prime}.

We consider a set \mathcal{R} of DD-subspaces of VV whose direct sum is V.V. We call \mathcal{R} a frame if all elements of \mathcal{R} are one dimensional. An element αNormD1(V)\alpha\in\operatorname{Norm}^{1}_{D}(V) is split by \mathcal{R} if α\alpha is the direct sum of the α|W,W.\alpha|_{W},\ W\in\mathcal{R}. If in addition \mathcal{R} is a frame and (wi)i(w_{i})_{i} is a DD-basis of VV consisting of elements of WW\bigcup_{W\in\mathcal{R}}W one calls (wi)i(w_{i})_{i} a splitting basis of α.\alpha.

Remark 3.2

[BT84b, 1.26] Every pair of DD-norms on VV has a common splitting basis.

Given two norms α,γNormD1(V)\alpha,\ \gamma\in\operatorname{Norm}^{1}_{D}(V) with common splitting basis (vi)(v_{i}) and a real number λ[0,1]\lambda\in[0,1] the convex combination of α\alpha with γ\gamma with λ\lambda is defined to be the DD-norm on VV split by (vi)(v_{i}) whose value at viv_{i} is

λα(vi)+(1λ)γ(vi).\lambda\alpha(v_{i})+(1-\lambda)\gamma(v_{i}).

This definition does not depend on the choice of the splitting basis and we get an affine structure on NormD1(V).\operatorname{Norm}^{1}_{D}(V). The AutD(V)\operatorname{Aut}_{D}(V)-action on NormD1(V),\operatorname{Norm}^{1}_{D}(V), more precisely

(g.α)(v):=α(g1v),(g.\alpha)(v):=\alpha(g^{-1}v),

restricts to a U(σ)\operatorname{U}(\sigma)-action on Normh1(V).\operatorname{Norm}^{1}_{h}(V).

The family of balls of a DD-norm α\alpha leads to the idea of a lattice function:

Λα(t):={vV|α(v)t},t.\Lambda_{\alpha}(t):=\{v\in V|\ \alpha(v)\geq t\},\ t\in\mbox{$\mathbb{R}$}.

Before we give the definition we want to remark that by an oDo_{D}-lattice in VV we mean a finitely generated oDo_{D}-submodule of VV which contains a DD-basis of V,V, i.e. we omit the word ”full”.

Definition 3.3

[BL02, I.2.1] A family (Λ(r))r(\Lambda(r))_{r\in\mbox{$\mathbb{R}$}} of oDo_{D}-lattices in VV is said to be an oDo_{D}-lattice function in VV if

  1. 1.

    Λ(r)Λ(s),\Lambda(r)\supseteq\Lambda(s),

  2. 2.

    Λ(r)=t<rΛ(t)\Lambda(r)=\bigcap_{t<r}\Lambda(t) and

  3. 3.

    Λ(r)𝔭D=Λ(r+1d)\Lambda(r)\mathfrak{p}_{D}=\Lambda(r+\frac{1}{d}),

for all r,sr,s\in\mbox{$\mathbb{R}$} with r<s.r<s. The set of oDo_{D}-lattice functions in VV is denoted by LattoD1(V).\operatorname{Latt}^{1}_{o_{D}}(V). For the right limit of Λ\Lambda at ss we write Λ(s+),\Lambda(s+), i.e.

Λ(s+):=t>sΛ(s).\Lambda(s+):=\bigcup_{t>s}\Lambda(s).

For tt\in\mbox{$\mathbb{R}$} we define t\lceil t\rceil to be smallest integer not smaller than t.t.

Proposition 3.4

[BL02, I.2.4] The map

αΛα\alpha\mapsto\Lambda_{\alpha}

is a bijection from NormD1(V)\operatorname{Norm}^{1}_{D}(V) to LattoD1(V).\operatorname{Latt}^{1}_{o_{D}}(V). Its inverse is given by

αΛ(v):=sup{t|vΛ(t)}.\alpha_{\Lambda}(v):=\sup\{t\in\mbox{$\mathbb{R}$}|\ v\in\Lambda(t)\}.

All notions for norms carry over to lattice functions in the following way. The dual of a lattice function Λ\Lambda is the lattice function which corresponds to αΛ#.\alpha^{\#}_{\Lambda}. A lattice function is called split by a given basis if the correspoding norm is split by this basis. The AutD(V)\operatorname{Aut}_{D}(V)-action on NormD1(V)\operatorname{Norm}^{1}_{D}(V) defines an AutD(V)\operatorname{Aut}_{D}(V)-action on LattoD1(V)\operatorname{Latt}^{1}_{o_{D}}(V) via push forward.

Proposition 3.5

[BS09, 3.3],[BL02, I.2.4] Let Λ\Lambda be an oDo_{D}-lattice function in V.V. Let gg be an element of AutD(V)\operatorname{Aut}_{D}(V) and let \mathcal{R} be a set of DD-subspaces of VV whose direct sum is V.V.

  1. 1.

    The function Λ\Lambda is split by \mathcal{R} if and only if, for all real numbers tt, the lattice Λ(t)\Lambda(t) is the direct sum of the oDo_{D}-modules WΛ(t),W\cap\Lambda(t), W.W\in\mathcal{R}.

  2. 2.

    If (vi)i(v_{i})_{i} is a splitting basis of Λ,\Lambda, then there are real numbers ai, 1im,a_{i},\ 1\leq i\leq m, such that, for all tt, we have

    Λ(t)=ivi𝔭D(tai)d.\Lambda(t)=\bigoplus_{i}v_{i}\mathfrak{p}_{D}^{\lceil(t-a_{i})d\rceil}.

    We call (ai)i(a_{i})_{i} the coordinate tuple of Λ\Lambda with respect to (vi).(v_{i}). The map which assigns the coordinate tuple to a lattice function split by (vi)(v_{i}) is an affine bijection onto m.\mbox{$\mathbb{R}$}^{m}.

  3. 3.

    We have (g.Λ)(t)=g(Λ(t)).(g.\Lambda)(t)=g(\Lambda(t)).

  4. 4.

    The dual Λ#\Lambda^{\#} of Λ\Lambda with respect to hh is the oDo_{D}-lattice function whose value at tt\in\mbox{$\mathbb{R}$} is

    {vV|h(v,Λ((t)+))𝔭D}.\{v\in V|\ h(v,\Lambda((-t)+))\subseteq\mathfrak{p}_{D}\}.
Proof.

All statements except for 4. can be found in part I of [BL02]. Point 4. has been proven in [BS09, 3.3] for the case D=k,D=k, but the proof is valid for the general case if one replaces oFo_{F} by oD.o_{D}.

The set of self-dual oDo_{D}-lattice functions is denoted by Latth1(V).\operatorname{Latt}^{1}_{h}(V). We recall that we have fixed an ϵ\epsilon-hermitian form hh on V.V.

Theorem 3.6

[BT87, 2.12] There is a unique affine and GG-equivariant bijection from 𝔅1(G)\mathfrak{B}^{1}(G) to Normh1(V).\operatorname{Norm}^{1}_{h}(V).

Propositions 3.4 and 3.5 imply the existence of a unique affine and GG-equivariant bijection from 𝔅1(G)\mathfrak{B}^{1}(G) to Latth1(V).\operatorname{Latt}^{1}_{h}(V). It defines on Latth1(V)\operatorname{Latt}^{1}_{h}(V) a system of apartments which correspond to the Witt-decompositions of V.V.

Definition 3.7

A set of one dimensional hh-isotropic DD-subspaces WlW_{l} of V,lL,V,\ l\in L, is said to be a Witt decomposition of VV if:

  1. 1.

    for every lLl\in L there is exactly one index lLl^{\prime}\in L such that h(Wl,Wl)h(W_{l},W_{l^{\prime}}) is non-zero,

  2. 2.

    the sum of the WlW_{l} is direct, and

  3. 3.

    the orthogonal complement WW of the sum of the WlW_{l} is anisotropic.

A lattice function is said to be split by a Witt decomposition {Wl|lL}\{W_{l}|\ l\in L\} if it is split by {Wl|lL}{W}.\{W_{l}|\ l\in L\}\cup\{W\}. A DD-basis of VV is adapted to {Wl|lL}\{W_{l}|\ l\in L\} if all basis elements lie in the union of WW and all Wl.W_{l}. Further we assume 0L0\not\in L and denote WW by W0.W_{0}.

Witt decompositions always exists, and even more, for every element of Latth1(V)\operatorname{Latt}^{1}_{h}(V) there is a splitting Witt decomposition. A proof for the latter in terms of norms can be found in [BT87, 2.13].

Remark 3.8

[BT87, 2.9]

  1. 1.

    The system of apartments of Latth1(V)\operatorname{Latt}^{1}_{h}(V) is the system of sets

    Latth,𝒮1(V):={ΛLatth1(V)|Λ is split by 𝒮},\operatorname{Latt}^{1}_{h,\mathcal{S}}(V):=\{\Lambda\in\operatorname{Latt}^{1}_{h}(V)|\ \Lambda\text{ is split by }\mathcal{S}\},

    where 𝒮\mathcal{S} runs over the set of Witt decompositions of VV with respect to h.h.

  2. 2.

    If a self dual DD-norm α\alpha split by a Witt decomposition {Wl|lL}\{W_{l}|\ l\in L\} satisfies

    α(w0)=12ν(h(w0,w0)),\alpha(w_{0})=\frac{1}{2}\nu(h(w_{0},w_{0})),

    for all w0W0w_{0}\in W_{0} (see [BT87, 2.9]).

Proposition 3.9

If α\alpha is a self dual DD-norm on VV and {Wl|lL}\{W_{l}|\ l\in L\} is a Witt decomposition splitting α,\alpha, then every orthogonal DD-basis of W0W_{0} splits α|W0\alpha|_{W_{0}}.

Proof.

Let ww^{\prime} and w′′w^{\prime\prime} be elements of W0W_{0} which are orthogonal to each other, then we have

α(w)=α(12(w+w′′)+12(ww′′))inf{α(12(w+w′′)),α(12(ww′′))}=α(w+w′′),\alpha\left(w^{\prime}\right)=\alpha\left(\tfrac{1}{2}\left(w^{\prime}+w^{\prime\prime}\right)+\tfrac{1}{2}\left(w^{\prime}-w^{\prime\prime}\right)\right)\geq\inf\{\alpha\left(\tfrac{1}{2}\left(w^{\prime}+w^{\prime\prime}\right)\right),\alpha\left(\tfrac{1}{2}\left(w^{\prime}-w^{\prime\prime}\right)\right)\}=\alpha\left(w^{\prime}+w^{\prime\prime}\right),

and the last equality follows from

h(12(w+w′′),12(w+w′′))=h(12(ww′′),12(ww′′))h\left(\tfrac{1}{2}\left(w^{\prime}+w^{\prime\prime}\right),\tfrac{1}{2}\left(w^{\prime}+w^{\prime\prime}\right)\right)=h\left(\tfrac{1}{2}\left(w^{\prime}-w^{\prime\prime}\right),\tfrac{1}{2}\left(w^{\prime}-w^{\prime\prime}\right)\right)

and ν(2)=0.\nu(2)=0.

4 The Moy–Prasad filtrations

An oDo_{D}-lattice function Λ\Lambda gives rise to an oko_{k}-lattice function in A:=EndD(V)A:=\operatorname{End}_{D}(V)

g~Λ(t):={aA|a(Λ(s))Λ(s+t),s}\tilde{g}_{\Lambda}(t):=\{a\in A|\ a(\Lambda(s))\subseteq\Lambda(s+t),\ \forall s\in\mbox{$\mathbb{R}$}\}

called the square lattice function of Λ\Lambda in A.A. It defines a Lie algebra filtration of Λ\Lambda in 𝔤\mathfrak{g} by intersection

𝔤Λ(t):=g~Λ(t)𝔤.\mathfrak{g}_{\Lambda}(t):=\tilde{g}_{\Lambda}(t)\cap\mathfrak{g}.

In [MP94] Moy–Prasad attached to every point xx of 𝔅1(G,k)\mathfrak{B}^{1}(G^{\prime},k^{\prime}) a filtration (𝔞x,t)t(\mathfrak{a}_{x,t})_{t\in\mathbb{R}} of Lie(G)(k)\operatorname{Lie}(G^{\prime})(k^{\prime}) for a reductive group GG^{\prime} defined over a non-Archimedean local field k.k^{\prime}. The next theorem states that in our special situation of unitary groups we do not need the quite involved description.

Theorem 4.1

[Lem09] The Lie algebra filtration of a point of 𝔅1(G)\mathfrak{B}^{1}(G) is exactly its Moy–Prasad filtration.

In this article the square lattice functions are the key for the rigidity of Lie algebra filtrations because of the following proposition.

Proposition 4.2

[BS09, 3.5] Let Λ\Lambda be an oDo_{D}-lattice function in V.V.

  1. 1.

    The oko_{k}-lattice function (σ(g~Λ(t)))t(\sigma(\tilde{g}_{\Lambda}(t)))_{t\in\mbox{$\mathbb{R}$}} is the square lattice of Λ#\Lambda^{\#} and we denote it by g~Λσ.\tilde{g}_{\Lambda}^{\sigma}. Square lattice functions fixed under σ\sigma are said to be self-dual.

  2. 2.

    The map

    ΛLatth1(V)g~Λ\Lambda\in\operatorname{Latt}^{1}_{h}(V)\mapsto\tilde{g}_{\Lambda}

    is injective and onto to the set Lattσ2(A)\operatorname{Latt}^{2}_{\sigma}(A) of self-dual square lattice functions.

Proof.

The proof in [BS09, 3.5] is valid if one replaces FF by D.D.

5 The centralizer

In the following the symbol Z(?)\operatorname{Z}_{*}(?) denotes the centralizer of ? in *. Let us fix an element βSkew(EndD(V),σ),\beta\in\operatorname{Skew}(\operatorname{End}_{D}(V),\sigma), such that the kk-algebra k[β]k[\beta] is a product of fields Ei,E_{i}, iI,i\in I, with identity element 1i.1_{i}. We call β\beta separable if Ei|1ikE_{i}|1_{i}k is separable for all i.i. The centralizers ZG(β)\operatorname{Z}_{G}(\beta) and Z𝐆(β)\operatorname{Z}_{\operatorname{\mathbf{G}}}(\beta) are denoted by HH and 𝐇.\operatorname{\mathbf{H}}. If β\beta is separable then the algebraic group 𝐇\operatorname{\mathbf{H}} is reductive and defined over k0k_{0} and its set of k0k_{0}-rational points is HH (see appendix A).

Notation 5.1

There is an action of σ\sigma on II via σ(1i)=:1σ(i).\sigma(1_{i})=:1_{\sigma(i)}. We denote by I0I_{0} the fixed point set of the action of σ\sigma on I,I, and we divide the set II0I\setminus I_{0} into two disjoint parts, i.e. we choose a positive part I+I_{+} and a negative part I,I_{-}, such that

σ(I+)=I.\sigma(I_{+})=I_{-}.

We define i:=σ(i)-i:=\sigma(i) and put Vi:=1iV.V_{i}:=1_{i}V. For iI0i\in I_{0} we denote by (Ei)0(E_{i})_{0} the set of fixed points of σ\sigma in Ei,E_{i}, and for iI+i\in I_{+} we put (Ei)0:=Ei.(E_{i})_{0}:=E_{i}.

The group HH is the product of sets of rational points of classical groups over the (Ei)0.(E_{i})_{0}. More precisely, for iI,i\in I, the EiE_{i}-algebra EndEikD(Vi)\operatorname{End}_{{E_{i}}\otimes_{k}{D}}(V_{i}) is EiE_{i}-algebra isomorphic to EndDi(Vi),\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i}), for some finite-dimensional vector space ViV^{\prime}_{i} over some skew field DiD^{\prime}_{i} central and of finite index over Ei.E_{i}. We define

𝐇i:={𝐔(σ|EndDi(Vi)),iI0𝐆𝐋Di(Vi),iI+\operatorname{\mathbf{H}}_{i}:=\left\{\begin{array}[]{cc}\operatorname{\mathbf{U}}(\sigma|_{\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i})}),&i\in I_{0}\\ \operatorname{\mathbf{GL}}_{D^{\prime}_{i}}(V^{\prime}_{i}),&i\in I_{+}\\ \end{array}\right.

and Hi:=𝐇i((Ei)0).H_{i}:=\operatorname{\mathbf{H}}_{i}((E_{i})_{0}). There is a canonical group isomorphism from HH to iI+I0Hi\prod_{i\in I_{+}\cup I_{0}}H_{i} which motivates the following definition for the Bruhat–Tits building of H:H:

𝔅1(H):=iI+I0𝔅1(𝐇i,(Ei)0).\mathfrak{B}^{1}(H):=\prod_{i\in I_{+}\cup I_{0}}\mathfrak{B}^{1}(\operatorname{\mathbf{H}}_{i},(E_{i})_{0}).
Remark 5.2

If β\beta is separable the above isomorphism extends to a k0k_{0}-isomorphism

𝐇iI0I+Res(Ei)0|k0(𝐇i)(k0).\displaystyle\operatorname{\mathbf{H}}\cong\prod_{i\in I_{0}\cup I_{+}}\operatorname{Res}_{(E_{i})_{0}|k_{0}}(\operatorname{\mathbf{H}}_{i})(k_{0}). (1)

Thus there is a isomorphism of affine buildings from 𝔅1(𝐇,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{H}},k_{0}) to 𝔅1(H).\mathfrak{B}^{1}(H).

The Lie algebra of HH is isomorphic to

iI0Skew(EndEikD(Vi),σ)iI+EndEikD(Vi)\bigoplus_{i\in I_{0}}\operatorname{Skew}(\operatorname{End}_{{E_{i}}\otimes_{k}{D}}(V_{i}),\sigma)\oplus\bigoplus_{i\in I_{+}}\operatorname{End}_{{E_{i}}\otimes_{k}{D}}(V_{i})

and we denote the iith factor by 𝔤i.\mathfrak{g}_{i}. We define the Lie algebra filtration of x=(xi)iI0I+x=(x_{i})_{i\in I_{0}\cup I_{+}} to be the direct sum of the Lie algebra filtrations of the xi,x_{i}, i.e.

𝔥x(t):=iI0I+(𝔤i)xi(t).\mathfrak{h}_{x}(t):=\bigoplus_{i\in I_{0}\cup I_{+}}(\mathfrak{g}_{i})_{x_{i}}(t).

where we take the square lattice function for iI+i\in I_{+} and identify EndEikD(Vi)\operatorname{End}_{{E_{i}}\otimes_{k}{D}}(V_{i}) with EndDi(Vi).\operatorname{End}_{D^{\prime}_{i}}(V_{i}).

6 Compatibility with the Lie algebra filtrations

Definition 6.1
  1. 1.

    A set with Lie algebra filtrations (LF-set) is a pair (X,L)(X,L) such that XX is a set, LL is a Lie algebra, and for each xXx\in X, there is Lie algebra filtration ((Lx)(t))t,((L_{x})(t))_{t\in\mbox{$\mathbb{R}$}}, i.e. a decreasing sequence of subsets of L.L.

  2. 2.

    Given two LF-sets (X,L)(X,L) and (X,L)(X^{\prime},L^{\prime}) and a map ϕ:LL\phi:L\mbox{$\rightarrow$}L^{\prime}

    1. (a)

      a map f:XXf:X\mbox{$\rightarrow$}X^{\prime} is compatible with the Lie algebra filtrations (CLF) if

      ϕ(Lx(t))=im(ϕ)Lf(x)(t)\phi(L_{x}(t))=\operatorname{im}(\phi)\cap L^{\prime}_{f(x)}(t)

      holds, for all xXx\in X and all real numbers tt.

    2. (b)

      g:XXg:X^{\prime}\mbox{$\rightarrow$}X is compatible with the Lie algebra filtrations (CLF) if

      ϕ(Lg(x)(t))=im(ϕ)Lx(t)\phi(L_{g(x^{\prime})}(t))=\operatorname{im}(\phi)\cap L^{\prime}_{x^{\prime}}(t)

      holds, for all xXx^{\prime}\in X^{\prime} and all real numbers tt.

In this article we consider buildings together with the set of rational points of a Lie algebra of an algebraic group. Mostly ϕ\phi is a canonical inclusion. An example of a CLF-map is given in [BL02]. The theorems in this section are valid without the assumption on the residue characteristic and the existence of an involution on D.D. Firstly, we introduce two important LF-sets after the following theorem.

Definition 6.2
  1. 1.

    Two oDo_{D}-lattice functions Λ\Lambda and Λ\Lambda^{\prime} on VV are equivalent if there is a real number tt such that

    Λ(s)=Λ(st)=:(Λ+t)(s),\Lambda(s)=\Lambda^{\prime}(s-t)=:(\Lambda^{\prime}+t)(s),

    for all s.s\in\mbox{$\mathbb{R}$}. The set of equivalence classes of elements of LattoD1(V)\operatorname{Latt}^{1}_{o_{D}}(V) is denoted by LattoD(V).\operatorname{Latt}_{o_{D}}(V). A map ff on LattoD1(V)\operatorname{Latt}^{1}_{o_{D}}(V) of the form

    f(Λ)=Λ+tf(\Lambda)=\Lambda+t

    is called a translation. For a map g,g, such that fgf\circ g exists, we define

    g+t:=fg.g+t:=f\circ g.
  2. 2.

    An action of AutD(V)\operatorname{Aut}_{D}(V) on LattoD(V)\operatorname{Latt}_{o_{D}}(V) is given by

    a.[Λ]:=[a.Λ],where (a.Λ)(t):=a(Λ(t)).a.[\Lambda]:=[a.\Lambda],\ \text{where }(a.\Lambda)(t):=a(\Lambda(t)).
  3. 3.

    The affine structure on LattoD1(V)\operatorname{Latt}^{1}_{o_{D}}(V) induces an affine structure on LattoD(V).\operatorname{Latt}_{o_{D}}(V).

Let us consider the non-extended building 𝔅(AutD(V))\mathfrak{B}(\operatorname{Aut}_{D}(V)) and the extended building 𝔅1(AutD(V)).\mathfrak{B}^{1}(\operatorname{Aut}_{D}(V)).

Theorem 6.3

[BL02, I.2.4],[BT84b, 2.13],[BT84b, 2.11 (iii)] The extended building of AutD(V)\operatorname{Aut}_{D}(V) is in affine and AutD(V)\operatorname{Aut}_{D}(V)-equivariant bijection with the set LattoD1(V).\operatorname{Latt}^{1}_{o_{D}}(V). For two such AutD(V)\operatorname{Aut}_{D}(V)-equivariant affine bijections ff and gg, there is a real number tt, such that

f=g+t,f=g+t,

and the map ff induces via

[x][f(x)][x]\mapsto[f(x)]

a unique affine and AutD(V)\operatorname{Aut}_{D}(V)-equivariant bijection from 𝔅(AutD(V))\mathfrak{B}(\operatorname{Aut}_{D}(V)) to LattoD(V).\operatorname{Latt}_{o_{D}}(V).

Remark 6.4

For x𝔅1(AutD(V))x\in\mathfrak{B}^{1}(\operatorname{Aut}_{D}(V)), we attach to xx and to [x][x] the square lattice function of a correspoding oDo_{D}-lattice function. By the above theorem both LF-sets (𝔅1(AutD(V)),EndD(V))(\mathfrak{B}^{1}(\operatorname{Aut}_{D}(V)),\operatorname{End}_{D}(V)) and (𝔅(AutD(V)),EndD(V))(\mathfrak{B}(\operatorname{Aut}_{D}(V)),\operatorname{End}_{D}(V)) are well-defined, i.e. do not depend on the choices made. In (𝔅(AutD(V)),EndD(V)),(\mathfrak{B}(\operatorname{Aut}_{D}(V)),\operatorname{End}_{D}(V)), a point is uniquely determined by its Lie algebra filtration (see [BL02, I.4.5]), whereas in (𝔅1(AutD(V)),EndD(V))(\mathfrak{B}^{1}(\operatorname{Aut}_{D}(V)),\operatorname{End}_{D}(V)) two points with the same Lie algebra filtration are translates of each other.

Theorem 6.5

[BL02, II.1.1] Let E|kE|k be a field extension in EndD(V),\operatorname{End}_{D}(V), G~:=AutD(V)\tilde{G}:=\operatorname{Aut}_{D}(V) and H~:=ZG~(E).\tilde{H}:=\operatorname{Z}_{\tilde{G}}(E). We consider the non-extended buildings 𝔅(G~)\mathfrak{B}(\tilde{G}) and 𝔅(H~)\mathfrak{B}(\tilde{H}) as LF-sets as constructed in the above remark. Then there is a unique CLF-map jEj_{E} from 𝔅(G~)E×\mathfrak{B}(\tilde{G})^{E^{\times}} to 𝔅(H~).\mathfrak{B}(\tilde{H}). The map has the following properties.

  1. 1.

    It is bijective.

  2. 2.

    It is H~\tilde{H}-equivariant.

  3. 3.

    It is affine.

The map jE:=(jE)1j^{E}:=(j_{E})^{-1} is the unique map satisfying 2. and 3..

In the above setting the CLF-property of jEj^{E} implies uniqueness.

Theorem 6.6

[BS09, 10.3] Under the assumptions of Theorem 6.5, suppose that the points y𝔅(G~)y\in\mathfrak{B}(\tilde{G}) and x𝔅(H~)x\in\mathfrak{B}(\tilde{H}) satisfy

𝔤~yEndEkD(V)𝔥~x.\tilde{\mathfrak{g}}_{y}\cap\operatorname{End}_{{E}\otimes_{k}{D}}(V)\supseteq\tilde{\mathfrak{h}}_{x}.

Then jE(x)=y.j^{E}(x)=y.

Proof.

In [BS09, 10.3] this theorem was proven for the case where D=kD=k and EE is generated by one element. The proof did not use the second assumption, and it carries over to DkD\neq k without changes. ∎

The map jEj^{E} is induced from a map between the extended buildings. We fix an EE-algebra isomorphism from EndEkD(V)\operatorname{End}_{{E}\otimes_{k}{D}}(V) to EndD(V)\operatorname{End}_{D^{\prime}}(V^{\prime}) where DD^{\prime} is a central skew field over EE and VV^{\prime} is a right DD^{\prime}-vector space.

Theorem 6.7

[BL02, II.3.1,II.4],[BT84b, 2.11 (iii)] There is a bijective affine H~\tilde{H}-equivariant CLF-map

j~E:LattoD1(V)LattoD1(V)E×\tilde{j}^{E}:\ \operatorname{Latt}^{1}_{o_{D^{\prime}}}(V^{\prime})\mbox{$\rightarrow$}\operatorname{Latt}^{1}_{o_{D}}(V)^{E^{\times}}

such that

[j~E(Λ)]=jE([Λ])[\tilde{j}^{E}(\Lambda)]=j^{E}([\Lambda])

for all ΛLattoD1(V).\Lambda\in\operatorname{Latt}^{1}_{o_{D^{\prime}}}(V^{\prime}). The image is the set of oDo_{D}-lattice functions which are in addition oEo_{E}-lattice functions. For every other bijective affine H~\tilde{H}-equivariant map jj from LattoD1(V)\operatorname{Latt}^{1}_{o_{D^{\prime}}}(V^{\prime}) to LattoD1(V)E×\operatorname{Latt}^{1}_{o_{D}}(V)^{E^{\times}} the composition j1j~Ej^{-1}\circ\tilde{j}^{E} is a translation of LattoD1(V).\operatorname{Latt}^{1}_{o_{D^{\prime}}}(V^{\prime}).

Proof.

The existence of j~E\tilde{j}^{E} is stated in Lemma [BL02, II.3.1]. The affineness is proven in [BL02, II.4] for jE,j_{E}, but the proof actually shows that j~E\tilde{j}^{E} is affine. The H~\tilde{H}-equivariance follows from the formula given in [BL02, II.3.1]. The fact that the image of jEj^{E} is the set of E×E^{\times}-fixed points implies that im(j~E)\operatorname{im}(\tilde{j}^{E}) only consists of oEo_{E}-lattice functions and contains a representative for every element of im(jE).\operatorname{im}(j^{E}). Being E×E^{\times}-equivariant and convex, the image of j~E\tilde{j}^{E} must contain every oEo_{E}-lattice function of LattoD1(V).\operatorname{Latt}^{1}_{o_{D}}(V). The last assertion follows directly from [BT84b, 2.11 (iii)]. ∎

7 CLF-map from 𝔅1(H)\mathfrak{B}^{1}(H) to 𝔅1(G)\mathfrak{B}^{1}(G)

We are now returning to the situation of section 5. Before we state the first theorem we give a definition in analogy with the set of fixed points in the building.

Definition 7.1

An oEo_{E}-oDo_{D}-lattice function of VV is an oDo_{D}-lattice function in VV which splits under (Vi)(V_{i}) such that, for every ii, the function

tΛ(t)Vit\mapsto\Lambda(t)\cap V_{i}

is an oEio_{E_{i}}-lattice function in Vi.V_{i}. We denote the set of oEo_{E}-oDo_{D}-lattice functions by LattoE,oD1(V).\operatorname{Latt}^{1}_{o_{E},o_{D}}(V).

The next theorem has been proven for D=kD=k in [BS09].

Theorem 7.2

There is an injective, affine and HH-equivariant CLF-map

j:𝔅1(H)𝔅1(G)j:\ \mathfrak{B}^{1}(H)\mbox{$\rightarrow$}\mathfrak{B}^{1}(G)

whose image in terms of lattice functions is the set of self-dual oEo_{E}-oDo_{D}-lattice functions in V.V.

Proof.

The product decomposition H=iI+I0ZU(hi|(Vi+Vi)2)(βi+βi)H=\prod_{i\in I_{+}\cup I_{0}}\operatorname{Z}_{\operatorname{U}(h_{i}|_{(V_{i}+V_{-i})^{2}})}(\beta_{i}+\beta_{-i}) leads us to three steps:

  1. 1.

    the case where EE is a field;

  2. 2.

    the case where I+I_{+} has cardinality one;

  3. 3.

    the general case.

Step 1: We use the map jEj^{E} from Theorem 6.5 and the following diagram.

LattoD(V)jELattoD(V)𝔅1(H)𝔅1(G)\begin{array}[]{ccc}\operatorname{Latt}_{o_{D^{\prime}}}(V^{\prime})&\stackrel{{\scriptstyle j^{E}}}{{\mbox{$\rightarrow$}}}&\operatorname{Latt}_{o_{D}}(V)\\ \uparrow&&\uparrow\\ \mathfrak{B}^{1}(H)&&\mathfrak{B}^{1}(G)\\ \end{array}

We have to prove that, for x𝔅1(H)x\in\mathfrak{B}^{1}(H), the square lattice function of jE(x)j^{E}(x) is self-dual. The latter follows from Theorem 6.6 because the square lattice function of xx is self-dual and jEj^{E} is a CLF-map. The assertion about the image of jE|𝔅1(H)j^{E}|_{\mathfrak{B}^{1}(H)} follows from 6.5 and the CLF-property.

Step 2: To prove the lemma we denote the unique element in I+I_{+} by ii and we define

Λ#i(t):={vVi|h(v,Λ((t)+))𝔭D}\Lambda^{\#_{-i}}(t):=\{v\in V_{-i}|\ h(v,\Lambda((-t)+))\subseteq\mathfrak{p}_{D}\}

for ΛLattoD1(Vi).\Lambda\in\operatorname{Latt}^{1}_{o_{D}}(V_{i}). The map

gAutD(Vi)(g,0)+σ((g1,0))AutD(Vi)AutD(Vi)g\in\operatorname{Aut}_{D}(V_{i})\mapsto(g,0)+\sigma((g^{-1},0))\in\operatorname{Aut}_{D}(V_{i})\oplus\operatorname{Aut}_{D}(V_{-i}) (2)

defines a kk-embedding from 𝐆𝐋D(Vi)\operatorname{\mathbf{GL}}_{D}(V_{i}) to 𝐆\operatorname{\mathbf{G}} mapping ZGLD(Vi)(βi)\operatorname{Z}_{\text{GL}_{D}(V_{i})}(\beta_{i}) onto H.H. An injective, affine and AutD(Vi)\operatorname{Aut}_{D}(V_{i})-equivariant CLF-map from 𝔅1(𝐆𝐋D(Vi),k)\mathfrak{B}^{1}(\operatorname{\mathbf{GL}}_{D}(V_{i}),k) to 𝔅1(G)\mathfrak{B}^{1}(G) in terms of lattice functions is given by

ΛLattoD1(Vi)(ΛΛ#i)Latth1(V).\Lambda\in\operatorname{Latt}^{1}_{o_{D}}(V_{i})\mapsto(\Lambda\oplus\Lambda^{\#_{-i}})\in\operatorname{Latt}^{1}_{h}(V). (3)

Now we apply Theorem 6.7 to finish step 2.

Step 3: We take the direct sum of the maps constructed in the steps before. ∎

8 Factorization

In order to analyze the set of CLF-maps from 𝔅1(H)\mathfrak{B}^{1}(H) to 𝔅1(G)\mathfrak{B}^{1}(G) this section reduces the problem to the case Where I0I+I_{0}\cup I_{+} has cardinality one.

Lemma 8.1

There is at most one index iIi\in I such that βi=0\beta_{i}=0 and if such an index exists, it has to be in I0.I_{0}.

Proof.

Assume that βi\beta_{i} and βj\beta_{j} are zero for two different indices ii and j.j. Taking a polynomial PP with coefficients in kk such that 1i=P(β)1_{i}=P(\beta) we obtain firstly

1i=1i1i=1iP(β)=1iP(0)1_{i}=1_{i}1_{i}=1_{i}P(\beta)=1_{i}P(0)

and secondly

0=1i1j=P(β)1j=P(0)1j0=1_{i}1_{j}=P(\beta)1_{j}=P(0)1_{j}

which is a contradiction. The second assertion follows from βi=σ(βi)-\beta_{-i}=\sigma(\beta_{i}) and the uniqueness. ∎

Let us recall that the embedding of 𝔥\mathfrak{h} into 𝔤\mathfrak{g} is realized by mapping an element (ai)iI0I+(a_{i})_{i\in I_{0}\cup I_{+}} of
(iI0Skew(EndEikD(Vi),σ))×(iI+EndEikD(Vi))(\prod_{i\in I_{0}}\operatorname{Skew}(\operatorname{End}_{{E_{i}}\otimes_{k}{D}}(V_{i}),\sigma))\times(\prod_{i\in I_{+}}\operatorname{End}_{{E_{i}}\otimes_{k}{D}}(V_{i})) to (iI0ai)+(iI+(aiσ(ai))).(\sum_{i\in I_{0}}a_{i})+(\sum_{i\in I_{+}}(a_{i}-\sigma(a_{i}))).

Assumption 8.2

For the rest of the section we fix elements y𝔅1(G)y\in\mathfrak{B}^{1}(G) and x𝔅1(H)x\in\mathfrak{B}^{1}(H) such that

𝔤y𝔥=𝔥x.\mathfrak{g}_{y}\cap\mathfrak{h}=\mathfrak{h}_{x}.

The element xx is given by a tuple (xi)iI0I+.(x_{i})_{i\in I_{0}\cup I_{+}}. The set Lie(𝐇i)((Ei)0)\operatorname{Lie}(\operatorname{\mathbf{H}}_{i})((E_{i})_{0}) is denoted by 𝔥i.\mathfrak{h}_{i}. For iI0i\in I_{0}, we write 𝔥xi\mathfrak{h}_{x_{i}} and 𝔥~xi\tilde{\mathfrak{h}}_{x_{i}} for the Lie algebra filtration and the square lattice function of xi,x_{i}, respectively.

Lemma 8.3

The lattice function corresponding to yy splits under {Vi|iI}.\{V_{i}|\ i\in I\}.

Proof.

The assertion is equivalent to the fact that all idempotents 1i1_{i} are elements of 𝔤~y(0).\tilde{\mathfrak{g}}_{y}(0).

Case 1: We first consider an index iI+.i\in I_{+}. Then 1i1_{i} is an element of 𝔥xi(0)\mathfrak{h}_{x_{i}}(0) and thus 1i1i1_{i}-1_{-i} is an element of 𝔤y(0)\mathfrak{g}_{y}(0) by 8.2. Therefore

1i+1i=(1i1i)2𝔤~y(0).1_{i}+1_{-i}=(1_{i}-1_{-i})^{2}\in\tilde{\mathfrak{g}}_{y}(0).

Hence 1i1_{i} and 1i1_{-i} are elements of 𝔤~y(0),\tilde{\mathfrak{g}}_{y}(0), since 22 is invertible in ok.o_{k}.

Case 2: We take an index iI0i\in I_{0} and we assume that βi\beta_{i} is not zero. Since βi\beta_{i} is skew-symmetric and central, we have, for all tt\in\mbox{$\mathbb{R}$},

βi𝔥~xi(t)=𝔥~xi(t+ν(βi))\beta_{i}\tilde{\mathfrak{h}}_{x_{i}}(t)=\tilde{\mathfrak{h}}_{x_{i}}(t+\nu(\beta_{i}))

and

βi𝔥xi(t)=𝔥~xi(t+ν(βi)){w𝔥~iσ|𝔥~i(w)=w}.\beta_{i}\mathfrak{h}_{x_{i}}(t)=\tilde{\mathfrak{h}}_{x_{i}}(t+\nu(\beta_{i}))\cap\{w\in\tilde{\mathfrak{h}}_{i}\mid\ \sigma|_{\tilde{\mathfrak{h}}_{i}}(w)=w\}.

By the invertibility of 22 in oko_{k}, every element of 𝔥~(t)\tilde{\mathfrak{h}}(t) is a sum of a skew-symmetric and a symmetric element of 𝔥~xi(t),\tilde{\mathfrak{h}}_{x_{i}}(t), which implies

𝔥~xi(0)\displaystyle\tilde{\mathfrak{h}}_{x_{i}}(0) =𝔥xi(0)+βi𝔥xi(ν(βi))\displaystyle=\mathfrak{h}_{x_{i}}(0)+\beta_{i}\mathfrak{h}_{x_{i}}(-\nu(\beta_{i}))
𝔤y(0)+𝔤y(ν(βi))𝔤y(ν(βi))\displaystyle\subseteq\mathfrak{g}_{y}(0)+\mathfrak{g}_{y}(\nu(\beta_{i}))\mathfrak{g}_{y}(-\nu(\beta_{i}))
𝔤~y(0).\displaystyle\subseteq\tilde{\mathfrak{g}}_{y}(0).

Thus the iith idempotent 1i1_{i} is an element of 𝔤~y(0).\tilde{\mathfrak{g}}_{y}(0).

Case 3: If there is an index i0i_{0} such that βi0=0,\beta_{i_{0}}=0, it is unique by Lemma 8.1 and the two cases above imply

1i0=1ii01i𝔤~y(0).1_{i_{0}}=1-\sum_{i\neq i_{0}}1_{i}\in\tilde{\mathfrak{g}}_{y}(0).

The idea of the proof of case 2 is taken from [BS09, 11.2]. We define 𝐆i:=𝐔(h|(Vi+Vi)2),\operatorname{\mathbf{G}}_{i}:=\operatorname{\mathbf{U}}(h|_{(V_{i}+V_{-i})^{2}}), for iI0I+.i\in I_{0}\cup I_{+}.

Corollary 8.4

For non-negative indices there are elements yi𝔅1(𝐆i,k0),y_{i}\in\mathfrak{B}^{1}(\operatorname{\mathbf{G}}_{i},k_{0}), such that the direct sum of the lattice functions of the yiy_{i} is the lattice function of y.y.

Analogous to the definitions for GG we use 𝔤i\mathfrak{g}_{i} and 𝔤yi\mathfrak{g}_{y_{i}} for Lie(𝐆i)(k0)\operatorname{Lie}(\operatorname{\mathbf{G}}_{i})(k_{0}) and the Lie algebra filtration of yi,y_{i}, respectively.

Lemma 8.5

For all iI0I+i\in I_{0}\cup I_{+} we have

𝔥xi=𝔥i𝔤yi.\mathfrak{h}_{x_{i}}=\mathfrak{h}_{i}\cap\mathfrak{g}_{y_{i}}.
Proof.

For tt\in\mbox{$\mathbb{R}$} we have

i𝔥xi(t)\displaystyle\bigoplus_{i}\mathfrak{h}_{x_{i}}(t) =𝔥(t)\displaystyle=\mathfrak{h}(t)
=𝔤y(t)𝔥\displaystyle=\mathfrak{g}_{y}(t)\cap\mathfrak{h}
=(𝔤y(t)(i𝔤i))𝔥\displaystyle=(\mathfrak{g}_{y}(t)\cap(\bigoplus_{i}\mathfrak{g}_{i}))\cap\mathfrak{h}
=(i𝔤yi(t))(i𝔥i)\displaystyle=(\bigoplus_{i}\mathfrak{g}_{y_{i}}(t))\cap(\bigoplus_{i}\mathfrak{h}_{i})
=i(𝔤yi(t)𝔥i),\displaystyle=\bigoplus_{i}(\mathfrak{g}_{y_{i}}(t)\cap\mathfrak{h}_{i}),

where ii runs over I0I+,I_{0}\cup I_{+}, and we obtain the assertion. ∎

The last two lemmas lead to a factorization of any CLF-map. More precisely we prove the following proposition.

Proposition 8.6

Let ψI\psi_{I} denote the canonical map from 𝔅1(i𝐆i,k0)\mathfrak{B}^{1}(\prod_{i}\operatorname{\mathbf{G}}_{i},k_{0}) to 𝔅1(𝐆,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{G}},k_{0}) which maps a tuple self-dual lattice functions to its sum. Every CLF-map jj from 𝔅1(H)\mathfrak{B}^{1}(H) to 𝔅1(G)\mathfrak{B}^{1}(G) factors under ψI,\psi_{I}, i.e. there is a unique map

τ:𝔅1(H)𝔅1(i𝐆i,k0) such that j=ψIτ.\tau:\ \mathfrak{B}^{1}(H)\mbox{$\rightarrow$}\mathfrak{B}^{1}(\prod_{i}\operatorname{\mathbf{G}}_{i},k_{0})\text{ such that }j=\psi_{I}\circ\tau.

The map τ\tau is

  1. 1.

    a CLF-map,

  2. 2.

    affine if jj is affine, and

  3. 3.

    HH-equivariant if jj is HH-equivariant.

Proof.

The image of jj is contained in the image of the injective, affine and iI0I+𝐆i(k0)\prod_{i\in I_{0}\cup I_{+}}\operatorname{\mathbf{G}}_{i}(k_{0})-equivariant map ψI\psi_{I} by Corollary 8.4. The CLF-property of τ\tau is a consequence of Lemma 8.5. This proves the proposition. ∎

9 The case where I+I_{+} is empty

We denote the following algebraic group

{(λ00λ1),(0λλ10)|λk¯×}\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\lambda^{-1}\end{array}\right),\left(\begin{array}[]{cc}0&\lambda\\ \lambda^{-1}&0\end{array}\right)|\ \lambda\in\bar{k}^{\times}\right\}

by 𝐎2is.\operatorname{\mathbf{O}}^{is}_{2}. It is a form of 𝐎2\operatorname{\mathbf{O}}_{2} over the prime field of k.k.

Remark 9.1
  1. 1.

    The group 𝐔(h)\operatorname{\mathbf{U}}(h) is k0k_{0}-isomorphic to 𝐎2is\operatorname{\mathbf{O}}^{is}_{2} if and only if D=k=k0,D=k=k_{0}, VV is two-dimensional over kk and isotropic with respect to hh and σ\sigma-orthogonal.

  2. 2.

    The connected component of 𝐎2is\operatorname{\mathbf{O}}^{is}_{2} is kk-isomorphic to 𝐆m,\operatorname{\mathbf{G}}_{m}, implying that 𝔅1(𝐎2is,k)\mathfrak{B}^{1}(\operatorname{\mathbf{O}}^{is}_{2},k) is affinely isomorphic to .\mbox{$\mathbb{R}$}.

  3. 3.

    All points of 𝔅1(𝐎2is,k)\mathfrak{B}^{1}(\operatorname{\mathbf{O}}^{is}_{2},k) have the same Lie algebra filtration. Especially all of its affine endomorphisms are CLF-maps.

The remark forces us to exclude 𝐎2is\operatorname{\mathbf{O}}^{is}_{2} from the factors.

Theorem 9.2

There exists only one CLF-map from 𝔅1(H)\mathfrak{B}^{1}(H) to 𝔅1(G)\mathfrak{B}^{1}(G) if I+I_{+} is empty and no group 𝐇i\operatorname{\mathbf{H}}_{i} is (Ei)0(E_{i})_{0}-isomorphic to 𝐎2is.\operatorname{\mathbf{O}}^{is}_{2}.

For the proof we need the following operations on square matrices.

Definition 9.3

For a square matrix B=(bi,j)Mr(D)B=(b_{i,j})\in M_{r}(D) the matrix B~\tilde{B} is defined to be (br+1j,r+1i)i,j,(b_{r+1-j,r+1-i})_{i,j}, i.e. B~\tilde{B} is obtained from BB by a reflection along the antidiagonal. We define further

Bρ:=(ρ(bi,j))i,j.B^{\rho}:=(\rho(b_{i,j}))_{i,j}.
Proof.

Applying Lemma 8.5 we can assume that EE is a field.

Case 1: β\beta is not zero. Compare with [BS09, 11.2]. We fix an arbitrary CLF-map jj from 𝔅1(H)\mathfrak{B}^{1}(H) to 𝔅1(G)\mathfrak{B}^{1}(G) and an arbitrary element xx of 𝔅1(H).\mathfrak{B}^{1}(H). By the same argument as in case 2 of Lemma 8.3 we obtain

𝔥~x(t)𝔤~j(x)(t)\tilde{\mathfrak{h}}_{x}(t)\subseteq\tilde{\mathfrak{g}}_{j(x)}(t)

for all real numbers t.t. Theorem 6.6 implies the uniqueness.

Case 2: β\beta is zero. If σ\sigma is of the second type there is a skew-symmetric non-zero element β\beta^{\prime} in kk and we can replace β\beta by β\beta^{\prime} and apply case 1. Thus we only need to consider the case where σ\sigma is of the first kind. We fix a point y𝔅1(G)y\in\mathfrak{B}^{1}(G) and fix an apartment containing y.y. This apartment is determined by a Witt decomposition. We choose an adapted basis (wi)1jm(w_{i})_{1\leq j\leq m} such that the Gram matrix Gram(wi)(h)\operatorname{Gram}_{(w_{i})}(h) of the ϵ\epsilon-hermitian form hh has the form

(0M0ϵM0000N)\left(\begin{array}[]{ccc}0&M&0\\ \epsilon M&0&0\\ 0&0&N\\ \end{array}\right)

with M:=antidiag(1,,1)M:=\operatorname{antidiag}(1,\ldots,1) and a diagonal regular matrix N.N. From 3.9 we deduce that the self-dual oDo_{D}-lattice function Λ\Lambda corresponding to yy is split by this basis. It is thus described by its intersections with the lines wiD,w_{i}D, i.e. there are real numbers aia_{i} such that

Λ(t)=iwi𝔭D(tai)d.\Lambda(t)=\bigoplus_{i}w_{i}\mathfrak{p}_{D}^{\lceil(t-a_{i})d\rceil}.

Thus the square lattice function of yy in tt is

𝔤~y(t)=i,j𝔭D(t+ajai)dEi,j\tilde{\mathfrak{g}}_{y}(t)=\bigoplus_{i,j}\mathfrak{p}_{D}^{\lceil(t+a_{j}-a_{i})d\rceil}E_{i,j}

where Ei,jE_{i,j} denotes the matrix with a 1 in the intersection of the iith row and the jjth column and zeros everywhere else. See for example [BL02, I.4.5].

It is enough to show that 𝔤~y\tilde{\mathfrak{g}}_{y} is determined by the Lie algebra filtration 𝔤y.\mathfrak{g}_{y}. Indeed, a class of oDo_{D}-lattice functions contains at most one self-dual lattice function. Thus the self-dual square lattice function of a point of 𝔅1(G)\mathfrak{B}^{1}(G) determines the point uniquely.

The adjoint involution of hh

BBσ=Gram(wi)(h)1(Bρ)TGram(wi)(h)B\mapsto B^{\sigma}=\operatorname{Gram}_{(w_{i})}(h)^{-1}(B^{\rho})^{T}\operatorname{Gram}_{(w_{i})}(h)

on Mm(D)\operatorname{M}_{m}(D) has under (wi)1jm(w_{i})_{1\leq j\leq m} the form

(B1,1B1,2B1,3B2,1B2,2B2,3B3,1B3,2B3,3)(C~2,2ϵC~1,2ϵMC3,2TNϵC~2,1C~1,1MC3,1TNϵN1C2,3TMN1C1,3TMN1C3,3TN).\left(\begin{array}[]{ccc}B_{1,1}&B_{1,2}&B_{1,3}\\ B_{2,1}&B_{2,2}&B_{2,3}\\ B_{3,1}&B_{3,2}&B_{3,3}\\ \end{array}\right)\mapsto\left(\begin{array}[]{ccc}\tilde{C}_{2,2}&\epsilon\tilde{C}_{1,2}&\epsilon MC_{3,2}^{T}N\\ \epsilon\tilde{C}_{2,1}&\tilde{C}_{1,1}&MC_{3,1}^{T}N\\ \epsilon N^{-1}C_{2,3}^{T}M&N^{-1}C_{1,3}^{T}M&N^{-1}C_{3,3}^{T}N\\ \end{array}\right).

The matrices B1,1,B1,2,B2,1B_{1,1},B_{1,2},B_{2,1} and B2,2B_{2,2} are r×rr\times r-matrices and C:=BρC:=B^{\rho} where rr is the Witt index of h.h. By the above calculation we obtain that Ei,jσE_{i,j}^{\sigma} is +Ei,j,+E_{i,j}, Ei,j-E_{i,j} or λEu,l\lambda E_{u,l} with (i,j)(u,l)(i,j)\neq(u,l) for some λD×.\lambda\in D^{\times}. From the self-duality of End(Λ)\operatorname{End}(\Lambda) and since 2 is invertible in oko_{k} we get:

𝔤y(t)k(Ei,jEi,jσ)=𝔭kt+ajai(Ei,jEi,jσ).\mathfrak{g}_{y}(t)\cap k(E_{i,j}-E_{i,j}^{\sigma})=\mathfrak{p}_{k}^{\lceil t+a_{j}-a_{i}\rceil}(E_{i,j}-E_{i,j}^{\sigma}).

For the calculation see Lemma 9.4 below. Thus we can obtain the exponent ajaia_{j}-a_{i} from the knowledge of the Lie algebra filtration if Ei,jE_{i,j} is not fixed by σ.\sigma. We now consider two cases.

Case 2.1: We assume that there is an anisotropic part in the Witt decomposition, i.e. NN occurs. The matrix Ei,mE_{i,m} is fixed by σ\sigma if and only if ii equals m.m. Thus from the knowledge of the Lie algebra filtration we know all differences aiama_{i}-a_{m} for all indexes ii different from m,m, and thus by subtractions we know the differences aiaja_{i}-a_{j} for all ii and j.j.

Case 2.2: Now we assume that there is no anisotropic part in the Witt decomposition. If ϵ\epsilon is 1,-1, no Ei,jE_{i,j} is fixed and we can obtain the differences aiaja_{i}-a_{j} for all ii and j.j. As a consequence, we only have to consider the case where hh is hermitian and D=kD=k (see [BT87, 1.14]). Here the matrix Ei,jE_{i,j} is fixed by σ\sigma if and only if i+j=m+1.i+j=m+1. Thus we can determine all differences aiaja_{i}-a_{j} where i+jm+1.i+j\neq m+1. If mm is at least 44 for an index ii there is an index kik\neq i with i+km+1i+k\neq m+1 and we can obtain aiam+1ia_{i}-a_{m+1-i} if we add akam+1ia_{k}-a_{m+1-i} to aiak.a_{i}-a_{k}. If mm equals 22, then the group 𝐆\operatorname{\mathbf{G}} is kk-isomorphic to 𝐎2is\operatorname{\mathbf{O}}^{is}_{2} which is excluded by the assumption of the theorem. ∎

The idea of taking the root system of 𝐆\operatorname{\mathbf{G}} over kk for the non-unitary step of the last proof was given by P. Broussous. To complete the proof we need the following lemma.

Lemma 9.4

For all tt\in\mbox{$\mathbb{R}$} we have

𝔭Dtdk=𝔭kt.\mathfrak{p}_{D}^{\lceil td\rceil}\cap k=\mathfrak{p}_{k}^{\lceil t\rceil}.
Proof.

For an element xx of kk, we have x𝔭Dtdx\in\mathfrak{p}_{D}^{\lceil td\rceil} if and only if ν(x)tdd\nu(x)\geq\frac{\lceil td\rceil}{d}. There are integers ll and ss such that 1sd1\leq s\leq d and

td=ld+s{\lceil td\rceil}=ld+s

Thus t=l+1{\lceil t\rceil}=l+1 and we get that ν(x)tdd\nu(x)\geq\frac{\lceil td\rceil}{d} if and only if ν(x)t\nu(x)\geq{\lceil t\rceil}. The ”only if” follows from ν(x).\nu(x)\in\mbox{$\mathbb{Z}$}.

Remark 9.5
  1. 1.

    In particular the proof of Theorem 9.2 shows that, if 𝐇i\operatorname{\mathbf{H}}_{i} is (Ei)0(E_{i})_{0} isomorphic to 𝐎2is,\operatorname{\mathbf{O}}^{is}_{2}, then βi\beta_{i} has to be zero and σ\sigma is of the first kind.

  2. 2.

    For positive indices 𝐇i\operatorname{\mathbf{H}}_{i} is not isomorphic to 𝐎2\operatorname{\mathbf{O}}_{2} because the latter is not connected.

  3. 3.

    Let us assume that β\beta is separable. The above remarks and 9.1 imply, for iI0I+i\in I_{0}\cup I_{+}, that 𝐎2is\operatorname{\mathbf{O}}_{2}^{is} is k0k_{0}-isomorphic to Res(Ei)0|k0(𝐇i)\operatorname{Res}_{(E_{i})_{0}|k_{0}}(\operatorname{\mathbf{H}}_{i}) if and only if 𝐇i\operatorname{\mathbf{H}}_{i} is (Ei)0(E_{i})_{0}-isomorphic to 𝐎2is.\operatorname{\mathbf{O}}_{2}^{is}.

10 Rigidity of Euclidean buildings

To have an approach to a uniqueness statement if there are no restrictions on II we show that Euclidean buildings are rigid for functionals.

Definition 10.1
  1. 1.

    A set with affine structure is a pair (S,)(S,*) consisting of a non-empty set SS and a map

    :[0,1]×S×SS,*:[0,1]\times S\times S\mbox{$\rightarrow$}S,

    which we denote

    ts1+(1t)s2:=(t,s1,s2).ts_{1}+(1-t)s_{2}:=*(t,s_{1},s_{2}).
  2. 2.

    An affine functional ff on a set SS with affine structure is an affine map from SS to ,\mbox{$\mathbb{R}$}, i.e.

    f(tx+(1t)y)=tf(x)+(1t)f(y),f(tx+(1-t)y)=tf(x)+(1-t)f(y),

    for all t[0,1]t\in[0,1] and x,yS.x,y\in S.

Proposition 10.2

Let Ω\Omega be a thick Euclidean building and |Ω||\Omega| be its geometric realization. Then every affine functional aa on |Ω||\Omega| is constant.

For the definition of a thick Euclidean building and its geometric realization see [Bro89, VI.3].

Proof.

Let C1,C2C_{1},\ C_{2} and C3C_{3} be three pairwise different adjacent chambers having a common co-dimension 1 face S.S. We denote by PiP_{i} the unique vertex of CiC_{i} which is not a vertex of S.S. The line segment [P1,P2][P_{1},P_{2}] meets [P1,P3][P_{1},P_{3}] and [P2,P3][P_{2},P_{3}] in a point Q|S¯|.Q\in|\bar{S}|. This can be seen as follows. We are working in three different apartments simultaneously. If Δij\Delta_{ij} denotes an apartment containing CiC_{i} and Cj,C_{j}, for different ii and j,j, the affine isomorphism from |Δ12||\Delta_{12}| to |Δ13||\Delta_{13}| fixing |Δ12Δ13||\Delta_{12}\cap\Delta_{13}| sends [P1,P2][P_{1},P_{2}] to [P1,P3][P_{1},P_{3}] and thus the unique intersection point in [P1,P2]|C1¯||C2¯|[P_{1},P_{2}]\cap|\bar{C_{1}}|\cap|\bar{C_{2}}| lies on [P1,P3],[P_{1},P_{3}], and similarly on [P3,P2].[P_{3},P_{2}]. Without loss of generality assume that a(Q)a(Q) vanishes. If a(P1)a(P_{1}) is negative then a(P2)a(P_{2}) and a(P3)a(P_{3}) are positive by the affineness of a.a. Thus a(Q)a(Q) is positive since it lies on [P2,P3].[P_{2},P_{3}]. A contradiction. Using galleries we obtain that aa is constant on vertices of the same type. An apartment is affinely generated by its vertices of a fixed type. Thus aa is constant on every apartment and therefore on |Ω|,|\Omega|, since any two apartments are connected by a gallery. ∎

Proposition 10.3

If 𝐆\operatorname{\mathbf{G}} is not k0k_{0}-isomorphic to 𝐎2is\operatorname{\mathbf{O}}_{2}^{is}, then every affine functional on 𝔅1(G)\mathfrak{B}^{1}(G) is constant.

Proof.

Not being k0k_{0}-isomorphic to 𝐎2is,\operatorname{\mathbf{O}}_{2}^{is}, the unitary group 𝐔(h)\operatorname{\mathbf{U}}(h) has no k0k_{0}-rational characters on the connected component of the identity implying that the non-extended and the extended buildings are equal (see B.3). If 𝐆\operatorname{\mathbf{G}} is totally isotropic, then 𝔅1(G)\mathfrak{B}^{1}(G) is a point and otherwise it is the geometric realization of a thick Euclidean building. Now we apply Proposition 10.2. ∎

Proposition 10.4
  1. 1.

    A k×k^{\times}-invariant affine functional on 𝔅1(𝐆𝐋D(V),k),\mathfrak{B}^{1}(\operatorname{\mathbf{GL}}_{D}(V),k), i.e. on LattoD1(V),\operatorname{Latt}^{1}_{o_{D}}(V), is constant.

  2. 2.

    Every k×k^{\times}-invariant affine functional on 𝔅1(𝐎2is,k)\mathfrak{B}^{1}(\operatorname{\mathbf{O}}_{2}^{is},k) is constant.

Proof.
  1. 1.

    Every fiber of a k×k^{\times}-invariant affine functional on 𝔅1(𝐆𝐋D(V),k)\mathfrak{B}^{1}(\operatorname{\mathbf{GL}}_{D}(V),k) is a union of classes of oDo_{D}-lattice functions. It therefore factorizes to an affine functional on 𝔅(𝐆𝐋D(V),k).\mathfrak{B}(\operatorname{\mathbf{GL}}_{D}(V),k). Now we apply Proposition 10.2.

  2. 2.

    This follows from part 1, because 𝔅1(𝐎2is,k)\mathfrak{B}^{1}(\operatorname{\mathbf{O}}_{2}^{is},k) is isomorphic to 𝔅1(𝐆m,k)\mathfrak{B}^{1}(\operatorname{\mathbf{G}}_{m},k) via a k×k^{\times}-equivariant affine bijection.

11 The general case

We introduce the notion of a translation in order to understand the class of affine CLF-maps which satisfy a mild equivariance condition.

Definition 11.1
  1. 1.

    A translation of LattoD1(V)\operatorname{Latt}^{1}_{o_{D}}(V) is a map from LattoD1(V)\operatorname{Latt}^{1}_{o_{D}}(V) to itself of the form

    ΛΛ+s\Lambda\mapsto\Lambda+s

    where ss is a real number.

  2. 2.

    A translation of Latth1(V)\operatorname{Latt}^{1}_{h}(V) is the identity if 𝐆\operatorname{\mathbf{G}} is not k0k_{0}-isomorphic to 𝐎2is.\operatorname{\mathbf{O}}^{is}_{2}.

  3. 3.

    A translation of 𝔅1(𝐇,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{H}},k_{0}) is a product of translations of the factors.

Remark 11.2

A translation of 𝔅1(𝐎2is,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{O}}_{2}^{is},k_{0}) is included in the definition because (𝐎2is)0(\operatorname{\mathbf{O}}_{2}^{is})^{0} is k0k_{0}-isomorphic to 𝐆m,\operatorname{\mathbf{G}}_{m}, i.e. there is a natural bijection from 𝔅1(𝐎2is,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{O}}_{2}^{is},k_{0}) to Lattk01(k0).\operatorname{Latt}^{1}_{k_{0}}(k_{0}).

Let us fix a map jj from 𝔅1(H)\mathfrak{B}^{1}(H) to 𝔅1(G)\mathfrak{B}^{1}(G) constructed as in the proof of Theorem 7.2. In this section we are going to prove:

Theorem 11.3

If ϕ\phi is an affine and Z(i𝐇i0((Ei)0))\operatorname{Z}(\prod_{i}\operatorname{\mathbf{H}}_{i}^{0}((E_{i})_{0}))-equivariant CLF-map from 𝔅1(H)\mathfrak{B}^{1}(H) to 𝔅1(G)\mathfrak{B}^{1}(G) then j1ϕj^{-1}\circ\phi is a translation of 𝔅1(H).\mathfrak{B}^{1}(H). In terms of lattice functions, the image of ϕ\phi is the set of self-dual oEo_{E}-oDo_{D}-lattice functions in VV and ϕ\phi is i𝐇i0((Ei)0)\prod_{i}\operatorname{\mathbf{H}}_{i}^{0}((E_{i})_{0})-equivariant.

The composition of j1j^{-1} with ϕ\phi is possible by the following fact.

Proposition 11.4

The image of a CLF-map from 𝔅1(H)\mathfrak{B}^{1}(H) to 𝔅1(G)\mathfrak{B}^{1}(G) is a subset of the set of oEo_{E}-oDo_{D}-lattice functions.

Proof.

By Lemma 8.5 we can assume that

I0I+={i}.I_{0}\cup I_{+}=\{i\}.

Case 1: (β=0\beta=0) The field EE is kk and there is nothing to prove.

Case 2: (iI0i\in I_{0} and β0\beta\neq 0) There is only one CLF-map by Theorem 9.2 and it fulfills the assertion by Theorem 7.2.

Case 3: (iI+i\in I_{+}) We choose two arbitrary points y𝔅1(G)y\in\mathfrak{B}^{1}(G) and x𝔅1(H)x\in\mathfrak{B}^{1}(H) such that 𝔤y𝔥=𝔥x.\mathfrak{g}_{y}\cap\mathfrak{h}=\mathfrak{h}_{x}. The lattice function Λ\Lambda of yy splits under (Vi,Vi)(V_{i},V_{-i}) by corollary 8.4. By self-duality it is sufficient to prove that ΛVi\Lambda\cap V_{i} is an oEio_{E_{i}}-lattice function. The building

𝔅1(H)=𝔅1(𝐆𝐋EikD(Vi),Ei)\mathfrak{B}^{1}(H)=\mathfrak{B}^{1}(\operatorname{\mathbf{GL}}_{{E_{i}}\otimes_{k}{D}}(V_{i}),E_{i})

is identified with the set of lattice functions over a skew field whose center is Ei.E_{i}. Thus we get

  • (aaσ)𝔤y(0)(a-a^{\sigma})\in\mathfrak{g}_{y}(0) for all aoEi×,a\in o_{E_{i}}^{\times},

  • πEiπEiσ𝔤y(1e),\pi_{E_{i}}-\pi_{E_{i}}^{\sigma}\in\mathfrak{g}_{y}(\frac{1}{e}), and

  • πEi1(πEi1)σ𝔤y(1e),\pi_{E_{i}}^{-1}-(\pi_{E_{i}}^{-1})^{\sigma}\in\mathfrak{g}_{y}(-\frac{1}{e}),

where ee is the ramification index of Ei|kE_{i}|k and πEi\pi_{E_{i}} is a prime element of Ei.E_{i}. We conclude that 1iΛ1_{i}\Lambda is an oEio_{E_{i}}-lattice function. ∎

of Theorem 11.3.

Proposition 11.4 enables us to define

τ:=j1ϕ\tau:=j^{-1}\circ\phi

because the image of jj is the set of all oEo_{E}-oDo_{D}-lattice functions in V.V.

If we know that τ\tau is a translation, then all assertions of Theorem 11.3 hold: A translation is a bijection and we conclude that ϕ\phi and jj have the same image. The i𝐇i0((Ei)0)\prod_{i}\operatorname{\mathbf{H}}_{i}^{0}((E_{i})_{0})-equivariance of ϕ\phi follows because jj and τ\tau are i𝐇i0((Ei)0)\prod_{i}\operatorname{\mathbf{H}}_{i}^{0}((E_{i})_{0})-equivariant.

We denote the coordinates of τ\tau by τi,\tau_{i}, iI0I+.i\in I_{0}\cup I_{+}. We need two steps to show that τ\tau is a translation. Let us fix iI0I+.i\in I_{0}\cup I_{+}.

Step 1: We prove that the coordinate τi\tau_{i} only depends on xi.x_{i}.

Case 1.1: iI0,i\in I_{0}, such that 𝐎2is\operatorname{\mathbf{O}}_{2}^{is} is not (Ei)0(E_{i})_{0}-isomorphic to 𝐇i\operatorname{\mathbf{H}}_{i}. By Theorem 9.2 we have τi(x)=xi,\tau_{i}(x)=x_{i}, for all x𝔅1(H).x\in\mathfrak{B}^{1}(H).

Case 1.2: We assume that we have an index iIi\in I such that 𝐇i\operatorname{\mathbf{H}}_{i} is (Ei)0(E_{i})_{0}-isomorphic to 𝐎2is\operatorname{\mathbf{O}}_{2}^{is}, whose building is affinely isomorphic to \mathbb{R} by Remark 11.2. If we fix an index t(II+){i}t\in(I\cup I_{+})\setminus\{i\} and coordinates xlx_{l} for l(II+){t}l\in(I\cup I_{+})\setminus\{t\}, then the map

xtτi(x)x_{t}\mapsto\tau_{i}(x)

is constant by Proposition 10.3 or 10.4. Thus τi\tau_{i} does not depend on xt.x_{t}.

Case 1.3: iI+.i\in I_{+}. Let us fix x𝔅1(H)x\in\mathfrak{B}^{1}(H). The lattice functions Λτi(x)\Lambda_{\tau_{i}(x)} and Λxi\Lambda_{x_{i}} are equivalent by the CLF-property of τ\tau. We define ai(x)a_{i}(x) to be the real number such that

Λτi(x)=Λxi+ai(x).\Lambda_{\tau_{i}(x)}=\Lambda_{x_{i}}+a_{i}(x).

The map aia_{i} is affine, since τi\tau_{i} is. By an analogous argument as in case 1.2 we have that aia_{i} does not depend on the ttth coordinate, for t(I+I0){i}t\in(I_{+}\cup I_{0})\setminus\{i\}.

Step 1 allows us to define a map τ~i\tilde{\tau}_{i} from 𝔅1(𝐇i,(Ei)0)\mathfrak{B}^{1}(\operatorname{\mathbf{H}}_{i},(E_{i})_{0}) to itself by

τ~i(xi):=τi(x),x𝔅1(H).\tilde{\tau}_{i}(x_{i}):=\tau_{i}(x),\ x\in\mathfrak{B}^{1}(H).

Step 2: Here we show that τ~i\tilde{\tau}_{i} is a translation. We firstly consider an index iI0i\in I_{0} such that 𝐇i\operatorname{\mathbf{H}}_{i} is (Ei)0(E_{i})_{0}-isomorphic to 𝐎2is.\operatorname{\mathbf{O}}_{2}^{is}. In this case we have k=k0k=k_{0} and Ei=1ik.E_{i}=1_{i}k. We identify 𝔅1(𝐎2is,k)\mathfrak{B}^{1}(\operatorname{\mathbf{O}}_{2}^{is},k) with .\mbox{$\mathbb{R}$}. The 𝐒𝐎2is(k)\operatorname{\mathbf{SO}}_{2}^{is}(k)-equivariance of τi\tau_{i} gives

τ~i(xi+1)=τ~i(xi)+1.\tilde{\tau}_{i}(x_{i}+1)=\tilde{\tau}_{i}(x_{i})+1.

The affineness property implies that τ~i\tilde{\tau}_{i} is a translation. For iI+i\in I_{+} the map aia_{i} in case 3 of step 1 is an affine functional and the k×k^{\times}-equivariance of τi\tau_{i} implies the k×k^{\times}-invariance of ai,a_{i}, because one gets in terms of lattice functions

Λπk+ai(Λπk)\displaystyle\Lambda\pi_{k}+a_{i}(\Lambda\pi_{k}) =τ~i(Λπk)\displaystyle=\tilde{\tau}_{i}(\Lambda\pi_{k})
=(τ~i(Λ))πk\displaystyle=(\tilde{\tau}_{i}(\Lambda))\pi_{k}
=Λ+ai(Λ)1\displaystyle=\Lambda+a_{i}(\Lambda)-1
=Λπk+ai(Λ),\displaystyle=\Lambda\pi_{k}+a_{i}(\Lambda),

where πk\pi_{k} is a prime element of kk. Thus aia_{i} is constant by Proposition 10.4. ∎

12 Torality

In this section we want to prove that the map constructed in section 7 respects apartments and is toral if β\beta is separable.

Definition 12.1

A map

f:𝔅1(G1,k0)𝔅1(G2,k0)f:\ \mathfrak{B}^{1}(G_{1},k_{0})\mbox{$\rightarrow$}\mathfrak{B}^{1}(G_{2},k_{0})

between two extended buildings of reductive groups defined over k0k_{0} is called toral if, for each maximal k0k_{0}-split torus SS of G1G_{1}, there is a maximal k0k_{0}-split torus TT of G2G_{2} containing SS such that ff maps the apartment corresponding to SS into the apartment corresponding to T.T. An analogous definition applies to maps between non-extended buildings.

Maximal k0k_{0}-split tori of 𝐆\operatorname{\mathbf{G}} can be characterized in terms of Witt decompositions. Given a Witt decomposition {WllL}\{W_{l}\mid l\in L\} of VV, there is exactly one maximal k0k_{0}-split torus TT of 𝐆\operatorname{\mathbf{G}} which satisfies

Z𝐆(T)(k0)={g𝐆(k0)g(Wl)Wl,lL,(gidV)(W0){0}}.\operatorname{Z}_{\operatorname{\mathbf{G}}}(T)(k_{0})=\{g\in\operatorname{\mathbf{G}}(k_{0})\mid g(W_{l})\subseteq W_{l},\ l\in L,\ (g-\operatorname{id}_{V})(W_{0})\subseteq\{0\}\}.

We recall, that W0W_{0} is the orthogonal complement of the sum of all Wl.W_{l}. Every maximal k0k_{0}-split torus arises in this way, because they are conjugate to each other by elements of 𝐆(k0).\operatorname{\mathbf{G}}(k_{0}). The DD-vector spaces Wl,lL,W_{l},\ l\in L, are exactly the irreducible T(k0)T(k_{0})-invariant DD-subspaces of the orthogonal complement of

{vVt(v)=v,tT(k0)}.\{v\in V\mid\ t(v)=v,t\in T(k_{0})\}.

Analogously one shows that the set of maximal kk-split tori of 𝐆𝐋D(V)\operatorname{\mathbf{GL}}_{D}(V) is one-to-one correspondence with the set of all decompositions of VV into one-dimensional DD-subspaces.

Lemma 12.2

The map ψI\psi_{I} from 𝔅1(i𝐆i,k0)\mathfrak{B}^{1}(\prod_{i}\operatorname{\mathbf{G}}_{i},k_{0}) to 𝔅1(𝐆,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{G}},k_{0}) defined in Proposition 8.6 by

ψI((Λi)iI0I+):=iI0I+Λi\psi_{I}((\Lambda_{i})_{i\in I_{0}\cup I_{+}}):=\bigoplus_{i\in I_{0}\cup I_{+}}\Lambda_{i}

is toral.

Proof.

For iI0I+,i\in I_{0}\cup I_{+}, let SiS_{i} be a maximal k0k_{0}-split torus of 𝐆i\operatorname{\mathbf{G}}_{i} and {WlilLi}\{W_{l}^{i}\mid\ l\in L_{i}\} be the corresponding Witt decomposition of Vi+Vi.V_{i}+V_{-i}. Further let Δi\Delta_{i} be the apartment of SiS_{i} in 𝔅1(𝐆i,k0).\mathfrak{B}^{1}(\operatorname{\mathbf{G}}_{i},k_{0}). Let αi\alpha_{i} be an element of Δi\Delta_{i} seen as a self dual DD-norm on Vi+Vi.V_{i}+V_{-i}. By [BT87, 2.9] the norm

α:=iαi\alpha:=\sum_{i}\alpha_{i}

has the form

α(iw0i)=12infi(ν(h(w0i,w0i)))\alpha(\sum_{i}w_{0}^{i})=\frac{1}{2}\inf_{i}(\nu(h(w_{0}^{i},w_{0}^{i})))

on Va:=iW0i,V_{a}:=\sum_{i}W_{0}^{i}, i.e. this form does not depend on (αi)i.(\alpha_{i})_{i}. Let us take a splitting Witt decomposition {WlalLa}\{W^{a}_{l}\mid\ l\in L_{a}\} of VaV_{a} for the restriction of α\alpha to Va.V_{a}. Then the torus iSi\prod_{i}S_{i} is a subtorus of the maximal k0k_{0}-split torus T,T, which corresponds to the Witt decompostion

iI0I+{a}{WlilLi}\bigcup_{i\in I_{0}\cup I_{+}\cup\{a\}}\{W_{l}^{i}\mid l\in L_{i}\}

of V,V, and ψI\psi_{I} maps iΔi\prod_{i}\Delta_{i} into the apartment of T.T.

Proposition 12.3

The map jj constructed in the proof of Theorem 7.2 maps apartments into apartments. Furthermore, jj is toral if β\beta is separable.

Without loss of generality we may assume that I0I+={i}I_{0}\cup I_{+}=\{i\} by Lemma 12.2.

iI+i\in I_{+}.

The map j~Ei\tilde{j}^{E_{i}} of Theorem 6.7 from 𝔅1(H)\mathfrak{B}^{1}(H) to

𝔅1(𝐆𝐋D(Vi),k)=𝔅1(Resk|k0(𝐆𝐋D(Vi)),k0)\mathfrak{B}^{1}(\operatorname{\mathbf{GL}}_{D}(V_{i}),k)=\mathfrak{B}^{1}(\operatorname{Res}_{k|k_{0}}(\operatorname{\mathbf{GL}}_{D}(V_{i})),k_{0})

maps apartments into apartments and is toral if βi\beta_{i} is separable by [BL02, 5.1]. We prove that the canonical map ϕ\phi, see (3), from 𝔅1(Resk|k0(𝐆𝐋D(Vi)),k0)\mathfrak{B}^{1}(\operatorname{Res}_{k|k_{0}}(\operatorname{\mathbf{GL}}_{D}(V_{i})),k_{0}) to 𝔅1(𝐆,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{G}},k_{0}) defined by

ΛLattoD1(Vi)(ΛΛ#i)Latth1(V)\Lambda\in\operatorname{Latt}^{1}_{o_{D}}(V_{i})\mapsto(\Lambda\oplus\Lambda^{\#_{-i}})\in\operatorname{Latt}^{1}_{h}(V)

is toral. A maximal kk-split torus SS of 𝐆𝐋D(Vi)\operatorname{\mathbf{GL}}_{D}(V_{i}) corresponds to a decomposition of ViV_{i} in one-dimensional DD-subspaces, i.e. there is a decomposition Vi=lVi,lV_{i}=\bigoplus_{l}V_{i,l} such that

S(k)=Z({gGLD(Vi)|g(Vi,l)Vi,l for all l}).S(k)=\operatorname{Z}(\{g\in\text{GL}_{D}(V_{i})|\ g(V_{i,l})\subseteq V_{i,l}\text{ for all }l\}).

Let Vi,jV_{-i,j} be the subspace of ViV_{-i} dual to Vi,j,V_{i,j}, i.e.

Vi,j:={vVih(v,Vi,k)={0},kj},V_{-i,j}:=\{v\in V_{-i}\mid\ h(v,V_{i,k})=\{0\},k\neq j\},

and let TT be the torus given by the decomposition

V=l(Vi,lVi,l).V=\bigoplus_{l}(V_{i,l}\oplus V_{-i,l}).

Under the map (2), from AutD(Vi)\operatorname{Aut}_{D}(V_{i}) to AutD(Vi)AutD(Vi)\operatorname{Aut}_{D}(V_{i})\oplus\operatorname{Aut}_{D}(V_{-i}) defined by

gAutD(Vi)(g,0)+σ((g1,0))AutD(Vi)AutD(Vi)g\in\operatorname{Aut}_{D}(V_{i})\mapsto(g,0)+\sigma((g^{-1},0))\in\operatorname{Aut}_{D}(V_{i})\oplus\operatorname{Aut}_{D}(V_{-i})

the set S(k)S(k) is mapped into T(k)T(k) and under ϕ\phi the apartment of SS is mapped into the apartment of T.T. Let SS^{\prime} and TT^{\prime} be the maximal k0k_{0}-split sub-tori of Resk|k0(S)\operatorname{Res}_{k|k_{0}}(S) and Resk|k0(T)\operatorname{Res}_{k|k_{0}}(T) respectively. The set S(k0)S^{\prime}(k_{0}) is mapped into (T𝐆)(k0)(T^{\prime}\cap\operatorname{\mathbf{G}})(k_{0}) under (2). The image of ϕ\phi only consists of self-dual lattice functions. Hence ϕ\phi seen as a map from 𝔅1(Resk|k0(𝐆𝐋D(Vi)),k0)\mathfrak{B}^{1}(\operatorname{Res}_{k|k_{0}}(\operatorname{\mathbf{GL}}_{D}(V_{i})),k_{0}) to 𝔅1(𝐆,k0)\mathfrak{B}^{1}(\operatorname{\mathbf{G}},k_{0}) is toral. ∎

From now on we assume iI0.i\in I_{0}. Here we need the notion of tori adapted to a decomposition of V.V.

Definition 12.4

Assume we are given a decomposition

V=V+VV0V=V^{\prime+}\oplus V^{\prime-}\oplus V^{\prime 0}

such that V+V^{\prime+} and VV^{\prime-} are maximal totally isotropic and V+VV^{\prime+}\oplus V^{\prime-} is orthogonal to V0V^{\prime 0} with respect to h.h. A maximal k0k_{0}-split torus TT of 𝐆\operatorname{\mathbf{G}} is adapted to (V+,V,V0)(V^{\prime+},V^{\prime-},V^{\prime 0}) if there is a Witt decomposition {Wl|lL}\{W_{l}|\ l\in L\} corresponding to TT with anisotropic part V0V^{\prime 0} such that

l(WlV+)=V+ and l(WlV)=V.\bigoplus_{l}(W_{l}\cap V^{\prime+})=V^{\prime+}\text{ and }\bigoplus_{l}(W_{l}\cap V^{\prime-})=V^{\prime-}.

We say that an apartment of 𝔅1(G)\mathfrak{B}^{1}(G) is adapted to (V+,V,V0)(V^{\prime+},V^{\prime-},V^{\prime 0}) if every lattice function in this apartment is split by (V+,V,V0).(V^{\prime+},V^{\prime-},V^{\prime 0}).

iI0i\in I_{0}.

We have E=Ei.E=E_{i}. There are a central division algebra Δ\Delta over EE and a finite-dimensional right vector space WW such that EndEkD(V)\operatorname{End}_{{E}\otimes_{k}{D}}(V) is EE-algebra isomorphic to EndΔ(W).\operatorname{End}_{\Delta}(W). We identify the EE-algebras EndEkD(V)\operatorname{End}_{{E}\otimes_{k}{D}}(V) and EndΔ(W)\operatorname{End}_{\Delta}(W) via a fixed isomorphism and we fix a signed hermitian form hEh_{E} which corresponds to the restriction σE\sigma_{E} of σ\sigma to the EE-algebra EndΔ(W).\operatorname{End}_{\Delta}(W). Let rr be the Witt index of hE.h_{E}. We fix a decomposition

W=(W+W)W0W=(W^{+}\oplus W^{-})\oplus W^{0} (4)

such that W+W^{+} and WW^{-} are maximal isotropic subspaces of WW contained in the orthogonal complement of W0.W^{0}. Let e+,ee_{+},e_{-} and e0e_{0} be the projections to the vector spaces W+,WW^{+},W^{-} and W0W^{0} via the direct sum (4). We define

V+:=e+V,V:=eV and V0:=e0V.V^{+}:=e^{+}V,\ V^{-}:=e^{-}V\text{ and }V^{0}:=e^{0}V.

Consider the following diagram.

𝔅1(H)𝔅1(𝐔((hE)|W0×W0),E0)×𝔅1(𝐆𝐋Δ(W+),E)𝔅(𝐆𝐋Δ(W+),E)jα𝔅1(G)𝔅1(𝐔(h|V0×V0),k0)×𝔅1(𝐆𝐋D(V+),k)𝔅(𝐆𝐋D(V+),k)\begin{matrix}\mathfrak{B}^{1}(H)&\leftarrow&\mathfrak{B}^{1}(\operatorname{\mathbf{U}}((h_{E})|_{W^{0}\times W^{0}}),E_{0})\times\mathfrak{B}^{1}(\operatorname{\mathbf{GL}}_{\Delta}(W^{+}),E)&\mbox{$\rightarrow$}&\mathfrak{B}(\operatorname{\mathbf{GL}}_{\Delta}(W^{+}),E)\\ \downarrow j&&\downarrow\alpha&&\mbox{$\downarrow$}\\ \mathfrak{B}^{1}(G)&\leftarrow&\mathfrak{B}^{1}(\operatorname{\mathbf{U}}(h|_{V^{0}\times V^{0}}),k_{0})\times\mathfrak{B}^{1}(\operatorname{\mathbf{GL}}_{D}(V^{+}),k)&\mbox{$\rightarrow$}&\mathfrak{B}(\operatorname{\mathbf{GL}}_{D}(V^{+}),k)\\ \end{matrix}

where the vertical maps are induced by j.j. The right horizontal maps map a pair (x,Λ)(x,\Lambda) to the class of Λ.\Lambda. The right vertical arrow satisfies the CLF-property and its image only consists of E×E^{\times}-fixed points of 𝔅(𝐆𝐋D(V+),k),\mathfrak{B}(\operatorname{\mathbf{GL}}_{D}(V^{+}),k), both properties inherited from j.j. Thus the map in the right column is jE,j^{E}, i.e. the inverse of jE,j_{E}, because otherwise we could construct a CLF-map from 𝔅(𝐆𝐋D(V+),k)E×\mathfrak{B}(\operatorname{\mathbf{GL}}_{D}(V^{+}),k)^{E^{\times}} to 𝔅(𝐆𝐋Δ(W+),E)\mathfrak{B}(\operatorname{\mathbf{GL}}_{\Delta}(W^{+}),E) different from jE,j_{E}, but such a CLF-map is unique by [BL02, II.1.1.]. Now jEj^{E} maps apartments into apartments which implies that jj maps apartments adapted to (W+,W,W0)(W^{+},W^{-},W^{0}) into apartments adapted to (V+,V,V0).(V^{+},V^{-},V^{0}).

We now prove that jj is toral if E|kE|k is separable. Let us assume that E|kE|k is separable. This implies that the right column jEj^{E} is toral by [BL02, 5.1] which implies the torality of α\alpha because the only maximal E0E_{0}-split torus of the anisotropic group 𝐔((hE)|W0×W0)\operatorname{\mathbf{U}}((h_{E})|_{W^{0}\times W^{0}}) is the trivial group. The torality of α\alpha implies the torality of jj on tori adapted to (W+,W,W0).(W^{+},W^{-},W^{0}). Hence jj is toral because the triple (W+,W,W0)(W^{+},W^{-},W^{0}) was chosen arbitrarily. ∎

Appendix

Appendix A The centralizer of a separable Lie algebra element

We prove the representation of the centralizer as a product of general linear groups and Weil restrictions of unitary groups if β\beta is separable. We still rely on the notation from section 2. For more details about the Weil restriction we recommend [Wei82, 1.3.] and [KMRT98, 20.5.]. The definitions of DiD^{\prime}_{i} and ViV^{\prime}_{i} are given after notation 5.1.

Proposition A.1

Let us assume that β\beta is separable. The centralizer Z𝐆(β)\operatorname{Z}_{\operatorname{\mathbf{G}}}(\beta) is k0k_{0}-isomorphic to

iI0Res(Ei)0|k0(𝐔(σ|EndEikD(Vi)))×iI+Res(Ei)0|k0(𝐆𝐋Di(Vi));\prod_{i\in I_{0}}\operatorname{Res}_{(E_{i})_{0}|k_{0}}(\operatorname{\mathbf{U}}(\sigma|_{\operatorname{End}_{{E_{i}}\otimes_{k}{D}}(V_{i})}))\times\prod_{i\in I_{+}}\operatorname{Res}_{(E_{i})_{0}|k_{0}}(\operatorname{\mathbf{GL}}_{D^{\prime}_{i}}(V^{\prime}_{i}));

in particular it is reductive and defined over k0.k_{0}.

Before we come to the proof, we need some preparations on restriction of scalars.

Lemma A.2

Let DD^{\prime} be a skew field of finite index such that the center, denoted by E,E, is a non-Archimedean local field. Let VV^{\prime} be a finite-dimensional right DD^{\prime}-vector space. Assume further that σ\sigma^{\prime} is an involution on EndD(V)\operatorname{End}_{D^{\prime}}(V^{\prime}) whose set of fixed points in EE is E0.E_{0}. Let k0k_{0} be a subfield of E0E_{0} such that E0|k0E_{0}|k_{0} is separable and finite. Then 𝐔(σk0idΩ)\operatorname{\mathbf{U}}({\sigma^{\prime}}\otimes_{k_{0}}{\operatorname{id}_{\Omega}}) is k0k_{0}-isomorphic to ResE0|k0(𝐔(σ))\operatorname{Res}_{E_{0}|k_{0}}(\operatorname{\mathbf{U}}(\sigma^{\prime}))

Remark A.3

If 𝐕\operatorname{\mathbf{V}} is an affine variety defined over E0E_{0} there is an isomorphism

ResE0|k0(𝐕)γ𝐕γ,\operatorname{Res}_{E_{0}|k_{0}}(\operatorname{\mathbf{V}})\cong\prod_{\gamma}\operatorname{\mathbf{V}}^{\gamma},

defined over the normal hull of E0|k0E_{0}|k_{0}, where γ\gamma passes through the set Γ\Gamma of all field embeddings E0ΩE_{0}\hookrightarrow\Omega fixing k0k_{0}, and one obtains 𝐕γ\operatorname{\mathbf{V}}^{\gamma} from 𝐕\operatorname{\mathbf{V}} in the following way. We choose an automorphism γ¯\bar{\gamma} of Ω\Omega whose restriction to E0E_{0} is γ.\gamma. Let P1,,PlP_{1},\ldots,P_{l} be polynomials in nn variables with coefficients in E0E_{0} such that

𝐕:={xΩt|Pi(x)=0,i=1,,l},\operatorname{\mathbf{V}}:=\{x\in\Omega^{t}|\ P_{i}(x)=0,\ i=1,\ldots,l\},

for some t.t\in\mathbb{N}. One defines

𝐕γ:={(γ¯(xj))1jt|x𝐕}={xΩt|Piγ(x)=0,i=1,,l},\operatorname{\mathbf{V}}^{\gamma}:=\{(\bar{\gamma}(x_{j}))_{1\leq j\leq t}|\ x\in\operatorname{\mathbf{V}}\}=\{x\in\Omega^{t}|\ P_{i}^{\gamma}(x)=0,\ i=1,\ldots,l\},

where PiγP_{i}^{\gamma} is the polynomial obtained from PiP_{i} by applying γ\gamma to the coefficients.

of Lemma A.2.

The Ω\Omega-algebra EndD(V)k0Ω{\operatorname{End}_{D^{\prime}}(V^{\prime})}\otimes_{k_{0}}{\Omega} is canonically isomorphic to EndD(V)E0E0k0Ω{\operatorname{End}_{D^{\prime}}(V^{\prime})}\otimes_{E_{0}}{{E_{0}}\otimes_{k_{0}}{\Omega}} and E0k0Ω{E_{0}}\otimes_{k_{0}}{\Omega} is isomorphic to Ω[E0:k0]\Omega^{[E_{0}:k_{0}]} via

ek0w(γ(e)w)γΓ.{e}\otimes_{k_{0}}{w}\mapsto(\gamma(e)w)_{\gamma\in\Gamma}.

We denote by Ωγ\Omega^{\gamma} the left E0E_{0}-vector space Ω\Omega under the action

e.ω:=γ(e)ω.e.\omega:=\gamma(e)\omega.

Thus

EndD(V)E0E0k0ΩγEndD(V)E0Ωγ.{\operatorname{End}_{D^{\prime}}(V^{\prime})}\otimes_{E_{0}}{{E_{0}}\otimes_{k_{0}}{\Omega}}\cong\prod_{\gamma}{\operatorname{End}_{D^{\prime}}(V^{\prime})}\otimes_{E_{0}}{\Omega^{\gamma}}.

The fact that σ\sigma^{\prime} fixes E0E_{0} implies that σk0idΩ{\sigma^{\prime}}\otimes_{k_{0}}{\operatorname{id}_{\Omega}} is the product of involutions σE0idΩγ.{\sigma^{\prime}}\otimes_{E_{0}}{\operatorname{id}_{\Omega^{\gamma}}}. It is enough to prove

𝐔(σE0idΩγ)=𝐔(σE0idΩ)γ.\operatorname{\mathbf{U}}({\sigma^{\prime}}\otimes_{E_{0}}{\operatorname{id}_{\Omega^{\gamma}}})=\operatorname{\mathbf{U}}({\sigma^{\prime}}\otimes_{E_{0}}{\operatorname{id}_{\Omega}})^{\gamma}.

To show the latter, we fix γ\gamma and we choose an extension γ¯\bar{\gamma} to Ω.\Omega. The ring isomorphism

Φ:EndD(V)E0ΩEndD(V)E0Ωγ\Phi:\ {\operatorname{End}_{D^{\prime}}(V^{\prime})}\otimes_{E_{0}}{\Omega}\mbox{$\rightarrow$}{\operatorname{End}_{D^{\prime}}(V^{\prime})}\otimes_{E_{0}}{\Omega^{\gamma}}

sending fE0ω{f}\otimes_{E_{0}}{\omega} to fE0γ¯(ω){f}\otimes_{E_{0}}{\bar{\gamma}(\omega)} satisfies

Φ(σE0idΩ)=(σE0idΩγ)Φ.\Phi\circ({\sigma^{\prime}}\otimes_{E_{0}}{\operatorname{id}_{\Omega}})=({\sigma^{\prime}}\otimes_{E_{0}}{\operatorname{id}_{\Omega^{\gamma}}})\circ\Phi.

Thus an element gg of EndD(V)E0Ω{\operatorname{End}_{D^{\prime}}(V^{\prime})}\otimes_{E_{0}}{\Omega} lies in 𝐔(σE0idΩ)\operatorname{\mathbf{U}}({\sigma^{\prime}}\otimes_{E_{0}}{\operatorname{id}_{\Omega}}) if and only if Φ(g)\Phi(g) lies in 𝐔(σE0idΩγ)\operatorname{\mathbf{U}}({\sigma^{\prime}}\otimes_{E_{0}}{\operatorname{id}_{\Omega^{\gamma}}}), which proves the lemma. ∎

of Proposition A.1.

Without loss of generality we assume that I0I+={i}.I_{0}\cup I_{+}=\{i\}.

Case 1: I0={i}.I_{0}=\{i\}. We fix an algebraic closure Ω\Omega of E.E. We have

ZEndD(V)k0Ω(βk01)\displaystyle\operatorname{Z}_{{\operatorname{End}_{D}(V)}\otimes_{k_{0}}{\Omega}}({\beta}\otimes_{k_{0}}{1}) ZEndD(V)(β)k0Ω\displaystyle\cong\ {\operatorname{Z}_{\operatorname{End}_{D}(V)}(\beta)}\otimes_{k_{0}}{\Omega}
EndDi(Vi)k0Ω.\displaystyle\cong\ {\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i})}\otimes_{k_{0}}{\Omega}.

The involution defining 𝐔(σ)\operatorname{\mathbf{U}}(\sigma) on EndD(V)k0Ω{\operatorname{End}_{D}(V)}\otimes_{k_{0}}{\Omega} is σk0id.{\sigma}\otimes_{k_{0}}{\operatorname{id}}. The above Ω\Omega-algebra isomorphism defines an involution σik0id{\sigma^{\prime}_{i}}\otimes_{k_{0}}{\operatorname{id}} on the right side of the equation where σi\sigma^{\prime}_{i} is an involution on EndDi(Vi)\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i}) whose set of fixed points in EE is E0.E_{0}. By Lemma A.2 the group U(σik0id)\operatorname{U}({\sigma^{\prime}_{i}}\otimes_{k_{0}}{\operatorname{id}}) is the Weil restriction of 𝐔(σi)\operatorname{\mathbf{U}}(\sigma^{\prime}_{i}) from E0E_{0} to k0.k_{0}.

Case 2: I+={i}I_{+}=\{i\}. The algebra EE is a direct product of two fields EiE_{i} and EiE_{-i} and we define E0E_{0} to be the set of fixed points of σ\sigma in E.E. In the following part of the proof we use the bijections

EiE0Ei,eiei+σ(ei)σ(ei).E_{i}\mbox{$\rightarrow$}E_{0}\mbox{$\rightarrow$}E_{-i},\ e_{i}\mapsto e_{i}+\sigma(e_{i})\mapsto\sigma(e_{i}).

We fix an algebraic closure Ω\Omega of E0.E_{0}. As in the proofs of Lemma A.2 and case 1, we get the product decomposition

ZEndD(V)k0Ω(βk01)\displaystyle\operatorname{Z}_{{\operatorname{End}_{D}(V)}\otimes_{k_{0}}{\Omega}}({\beta}\otimes_{k_{0}}{1}) ZEndD(V)(β)k0Ω\displaystyle\cong\ {\operatorname{Z}_{\operatorname{End}_{D}(V)}(\beta)}\otimes_{k_{0}}{\Omega}
(EndDi(Vi)(EndDi(V))op)k0Ω\displaystyle\cong\ {(\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i})\oplus(\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}))^{op})}\otimes_{k_{0}}{\Omega}
γ((EndDi(Vi)(EndDi(Vi))op)E0Ωγ)\displaystyle\cong\ \prod_{\gamma}({(\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i})\oplus(\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i}))^{op})}\otimes_{E_{0}}{\Omega^{\gamma}})
γ((EndDi(Vi)(Ei)0Ωγ)(EndDi(Vi)(Ei)0Ωγ)op).\displaystyle\cong\ \prod_{\gamma}(({\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i})}\otimes_{(E_{i})_{0}}{\Omega^{\gamma}})\oplus({\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i})}\otimes_{(E_{-i})_{0}}{\Omega^{\gamma}})^{op}).

The unitary group

{g(EndDi(Vi)EiΩ)(EndDi(Vi)(Ei)0Ω)op|g(σE0idΩ)(g)=1}\{g\in({\operatorname{End}_{D^{\prime}_{i}}(V^{\prime}_{i})}\otimes_{E_{i}}{\Omega})\oplus({\operatorname{End}_{D^{\prime}_{-i}}(V^{\prime}_{-i})}\otimes_{(E_{-i})_{0}}{\Omega})^{op}|\ g({\sigma}\otimes_{E_{0}}{\operatorname{id}_{\Omega}})(g)=1\}

is EiE_{i}-isomorphic to 𝐆𝐋Di(Vi).\operatorname{\mathbf{GL}}_{D^{\prime}_{i}}(V^{\prime}_{i}). We conclude as in the proof of Lemma A.2. ∎

Appendix B Rational characters

In this section we show that the extended and the non-extended buildings of GG are equal in almost all cases. The following statement is common knowledge, but we give a prove for the sake of completeness. Let X(?)k0X^{*}(?)_{k_{0}} denote the set of k0k_{0}-rational characters of ??. Let us recall that by definition 𝔅1(G)\mathfrak{B}^{1}(G) is different from 𝔅(G)\mathfrak{B}(G) if and only if the set X(𝐆0)k0X^{*}(\operatorname{\mathbf{G}}^{0})_{k_{0}} is non-trivial (see [BT84a, 4.2.16]).

Definition B.1

The special unitary group 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) of hh is a connected reductive group defined over k0,k_{0}, whose set of rational points, which we denote by SU(h)\operatorname{SU}(h), is the intersection of 𝐔(h)(k0)\operatorname{\mathbf{U}}(h)(k_{0}) with the kernel of the reduced norm on EndD(V).\operatorname{End}_{D}(V).

Remark B.2

[PR94, 2.15]

  1. 1.

    If σ\sigma is a unitary involution, then 𝐆\operatorname{\mathbf{G}} is a kk-form of 𝐆𝐋md(k¯).\operatorname{\mathbf{GL}}_{md}(\bar{k}).

  2. 2.

    If σ\sigma is symplectic, then 𝐆\operatorname{\mathbf{G}} is kk-isomorphic to 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) which is a kk-form of the symplectic group 𝐒𝐩md(k¯),\operatorname{\mathbf{Sp}}_{md}(\bar{k}), in particluar 𝐆\operatorname{\mathbf{G}} is semisimple.

  3. 3.

    If σ\sigma is orthogonal, then 𝐆0\operatorname{\mathbf{G}}^{0} is kk-isomorphic to 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) which is a kk-form of the special orthogonal group 𝐒𝐎md(k¯),\operatorname{\mathbf{SO}}_{md}(\bar{k}), in particular 𝐆0\operatorname{\mathbf{G}}^{0} is semisimple if md2.md\neq 2.

We further need the following theorem.

Proposition B.3

X(𝐆0)k01X^{*}(\operatorname{\mathbf{G}}^{0})_{k_{0}}\neq 1 if and only if

m=2 and d=1 and σ is orthogonal and h is isotropic. m=2\text{ and }d=1\text{ and }\sigma\text{ is orthogonal and }h\text{ is isotropic. } (5)
Lemma B.4

Let L|LL|L^{\prime} be a field extension and n.n\in\mathbb{N}. Let L¯\bar{L} be an algebraic closure of L.L. Let ϕ\phi be an LL^{\prime}-algebra monomorphism from Mn(L)\operatorname{M}_{n}(L) into Mn(L¯).\operatorname{M}_{n}(\bar{L}). Then there is an element ψ\psi in Gal(L¯|L)\mathop{\text{Gal}}(\bar{L}|L^{\prime}) inducing an LL^{\prime}-algebra automorphism of Mn(L¯)\operatorname{M}_{n}(\bar{L}) via

Ψ((aij)ij):=(ψ(aij))ij,\Psi((a_{ij})_{ij}):=(\psi(a_{ij}))_{ij},

such that Ψϕ\Psi\circ\phi is an LL-algebra monomorphism from Mn(L)\operatorname{M}_{n}(L) into Mn(L¯).\operatorname{M}_{n}(\bar{L}).

Proof.

The map from Mn(L)LL¯{\operatorname{M}_{n}(L^{\prime})}\otimes_{L^{\prime}}{\bar{L}} to Mn(L¯)\operatorname{M}_{n}(\bar{L}) defined by

xLyϕ(x)y{x}\otimes_{L^{\prime}}{y}\mapsto\phi(x)y

is surjective, and thus im(ϕ)\operatorname{im}(\phi) contains a L¯\bar{L}-basis of Mn(L¯).\operatorname{M}_{n}(\bar{L}). In particular ϕ(L)\phi(L) is a subset of the center L¯\bar{L} of Mn(L¯).\operatorname{M}_{n}(\bar{L}). Now choose ψGal(L¯|L)\psi\in\mathop{\text{Gal}}(\bar{L}|L^{\prime}) such that ψ1|L\psi^{-1}|_{L} equals ϕ.\phi.

B.3.

A semisimple group is perfect and has therefore no characters. By remark B.2 the only cases left are:

  1. 1.

    σ\sigma is unitary.

  2. 2.

    σ\sigma is orthogonal and md=2md=2 and not (5).

  3. 3.

    Situation (5).

Case 1: We have D=kD=k by Remark 2.1. There is an isomorphism of k¯\bar{k}-algebras with involution

(Endk(V)k0k¯,σk0idk¯)(Mm(k¯)×Mm(k¯),σ~)({\operatorname{End}_{k}(V)}\otimes_{k_{0}}{\bar{k}},{\sigma}\otimes_{k_{0}}{\operatorname{id}_{\bar{k}}})\cong(\operatorname{M}_{m}(\bar{k})\times\operatorname{M}_{m}(\bar{k}),\tilde{\sigma})

with

σ~(B,C)=(CT,BT),\tilde{\sigma}(B,C)=(C^{T},B^{T}),

and we have k¯\bar{k}-isomorphisms

𝐆=U(σk0idk¯)U(σ~)GLm(k¯).\operatorname{\mathbf{G}}=\operatorname{U}({\sigma}\otimes_{k_{0}}{\operatorname{id}_{\bar{k}}})\cong\operatorname{U}(\tilde{\sigma})\cong\text{GL}_{m}(\bar{k}).

The last isomorphism is induced by the projection to the first coordinate. Let χ\chi be a k0k_{0}-rational character of 𝐆\operatorname{\mathbf{G}}. A character of GLm(k¯)\text{GL}_{m}(\bar{k}) is a power of the determinant. Thus, because of Lemma B.4, χ|G\chi|_{G} must be a power of det|G\det|_{G} or ρdet|G.\rho\circ\det|_{G}. We now fix a basis of VV to get a kk-isomorphism from Mm(k)\operatorname{M}_{m}(k) to Endk(V).\operatorname{End}_{k}(V). The involution σ()ρ\sigma\circ()^{\rho} is conjugate to the transpose map, which implies

χ(x)1=χ(σ(x))=ρ(χ(x))=χ(x),\chi(x)^{-1}=\chi(\sigma(x))=\rho(\chi(x))=\chi(x),

for all xG.x\in G. The last equality follows from χ(G)k0\chi(G)\subseteq k_{0}. In particular the only possible values of χ\chi on GG are 1 and 1.-1. Thus χ\chi is trivial, because 𝐆\operatorname{\mathbf{G}} is connected and GG is Zariski-dense in 𝐆\operatorname{\mathbf{G}} by [Bor91, 18.3].

Case 2: We have k=k0,k=k_{0}, since σ\sigma is orthogonal. If d=2d=2 There is an element aSU(h){1,1},a\in\operatorname{SU}(h)\setminus\{1,-1\}, because SU(h)\operatorname{SU}(h) is Zariski-dense in 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) by [Bor91, 18.3]. We have that k[a]k[a] is its own centralizer in D,D, because the index of DD is two, in particular the commutative group SU(h)\operatorname{SU}(h) is a subset of k[a]k[a]. if we introduce a kk-basis of DD which contains 11 and aa, then the identity from SU(h)\operatorname{SU}(h) to U(σ|k[a])\operatorname{U}(\sigma|_{k[a]}) can be extended to a kk-isomorphism from 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) to U(σ|k[a]).\operatorname{U}(\sigma|_{k[a]}). By case 1 there is no kk-rational character on 𝐒𝐔(h).\operatorname{\mathbf{SU}}(h).

Let us now assume d=1d=1 and 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) is anisotropic. There is a kk-basis of VV such that the Gram matrix of hh is of the form

(e00f),\left(\begin{array}[]{cc}e&0\\ 0&f\end{array}\right),

and we identify Endk(V)\operatorname{End}_{k}(V) with M2(k).\operatorname{M}_{2}(k). A short calculation shows that

𝐒𝐔(h)={(acfcea)a,ck¯s.t.a2+efc2=1}.\operatorname{\mathbf{SU}}(h)=\left\{\left(\begin{array}[]{cc}a&cf\\ -ce&a\end{array}\right)\mid a,c\in\bar{k}\ s.t.\ a^{2}+efc^{2}=1\right\}.

We fix square roots e\sqrt{e} and f.\sqrt{-f}. The conjugation with

(efe2f2)\left(\begin{array}[]{cc}\sqrt{e}&\sqrt{-f}\\ \frac{\sqrt{e}}{2}&-\frac{\sqrt{-f}}{2}\end{array}\right)

maps 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) to 𝐒𝐎2is\operatorname{\mathbf{SO}}_{2}^{is}. The explicit formula for the map is

(acfcea)(acef00a+cef),ef:=ef.\left(\begin{array}[]{cc}a&cf\\ -ce&a\end{array}\right)\mapsto\left(\begin{array}[]{cc}a-c\sqrt{-ef}&0\\ 0&a+c\sqrt{-ef}\end{array}\right),\ \sqrt{-ef}:=\sqrt{-e}\sqrt{f}.

Thus a character of 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) is of the form

(acfcea)(a+cef)z,\left(\begin{array}[]{cc}a&cf\\ -ce&a\end{array}\right)\mapsto(a+c\sqrt{-ef})^{z},

for some integer z.z. The inverse of (a+cef)(a+c\sqrt{-ef}) is (acef).(a-c\sqrt{-ef}). If zz is positive, we apply the binomial expansion to get coeffitients α\alpha und γ\gamma in kk such that

(a+cef)z=α+γef.(a+c\sqrt{-ef})^{z}=\alpha+\gamma\sqrt{-ef}.

The element γ\gamma is zero, because efk\sqrt{-ef}\notin k since hh is anisotropic. Thus a kk-rational character χ\chi of 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) satisfies

χ(x)=χ(x1),\chi(x)=\chi(x^{-1}),

for all xSU(h).x\in\operatorname{SU}(h). The density of SU(h)\operatorname{SU}(h) in 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) and the connectivity of 𝐒𝐔(h)\operatorname{\mathbf{SU}}(h) imply that χ\chi is trivial.

Case 3: Here 𝐆\operatorname{\mathbf{G}} is k0k_{0}-isomorphic to 𝐎2is\operatorname{\mathbf{O}}_{2}^{is} implying that 𝐆0\operatorname{\mathbf{G}}^{0} has non-trivial k0k_{0}-rational characters. ∎

References

  • [BK93] C.J. Bushnell and P.C. Kutzko. The admissible dual of GL(N){\rm GL}(N) via compact open subgroups, volume 129 of Ann. of Math. Studies. Princeton Univ. Press, Princeton, NJ, 1993.
  • [BL02] P. Broussous and B. Lemaire. Building of GL(m,D){\rm GL}(m,D) and centralizers. Transform. Groups, 7(1):15–50, 2002.
  • [Bor91] A. Borel. Linear algebraic groups, volume 126 of Grad. Texts in Math. Springer-Verlag, New York, 1991. 2nd enl. ed.
  • [Bro89] K.S. Brown. Buildings. Springer-Verlag, New York, 1989.
  • [BS09] P. Broussous and S. Stevens. Buildings of classical groups and centralizers of Lie algebra elements. J. Lie Theory, 19(1):55–78, 2009.
  • [BSS10] P. Broussous, V. Sécherre, and S. Stevens. Smooth representations of GL(m,D), V: Endo-classes. arXiv:1004.5032v1, 2010.
  • [BT84a] F. Bruhat and J. Tits. Groupes réductifs sur un corps local. II. Schémas en groupes. Existence d’une donnée radicielle valuée. Inst. Hautes Études Sci. Publ. Math., (60):197–376, 1984.
  • [BT84b] F. Bruhat and J. Tits. Schémas en groupes et immeubles des groupes classiques sur un corps local. Bull. Soc. Math. France, 112(2):259–301, 1984.
  • [BT87] F. Bruhat and J. Tits. Schémas en groupes et immeubles des groupes classiques sur un corps local. II. Groupes unitaires. Bull. Soc. Math. France, 115(2):141–195, 1987.
  • [KMRT98] M.-A. Knus, A. Merkurjev, M. Rost, and J.-P. Tignol. The book of involutions, volume 44 of AMS Colloquium Publications. Amer. Math. Soc., Providence, RI, 1998.
  • [Lan00] E. Landvogt. Some functorial properties of the Bruhat-Tits building. J. Reine Angew. Math., 518:213–241, 2000.
  • [Lem09] B. Lemaire. Comparison of lattice filtrations and Moy-Prasad filtrations for classical groups. J. Lie Theory, 19(1):29–54, 2009.
  • [MP94] A. Moy and G. Prasad. Unrefined minimal KK-types for pp-adic groups. Invent. Math., 116(1-3):393–408, 1994.
  • [PR94] V. Platonov and A. Rapinchuk. Algebraic groups and number theory, volume 139 of Pure and Appl. Math. Acad. press, Inc., BOSTON, 1994. Transl. from the 1991 Russ. orig. by R. Rowen.
  • [PY02] G. Prasad and J.-K. Yu. On finite group actions on reductive groups and buildings. Invent. Math., 147(3):545–560, 2002.
  • [Sch85] W. Scharlau. Quadratic and Hermitian Forms. Springer-Verlag, Berlin and Heidelberg, 1985.
  • [Séc04] V. Sécherre. Représentations lisses de GL(m,D){\rm GL}(m,D). I. Caractères simples. Bull. Soc. Math. France, 132(3):327–396, 2004.
  • [Séc05a] V. Sécherre. Représentations lisses de GL(m,D){\rm GL}(m,D). II. β\beta-extensions. Compos. Math., 141(6):1531–1550, 2005.
  • [Séc05b] V. Sécherre. Représentations lisses de GL(m,D){\rm GL}(m,D). III. Types simples. Ann. Sci. École Norm. Sup. (4), 38(6):951–977, 2005.
  • [SS08] V. Sécherre and S. Stevens. Représentations lisses de GLm(D){\rm GL}_{m}(D). IV. Représentations supercuspidales. J. Inst. Math. Jussieu, 7(3):527–574, 2008.
  • [SS11] V. Sécherre and S. Stevens. Smooth representations of GLm(D){\rm GL}_{m}(D) vi: Semisimple types. Int. Math. Res. Not., 2011.
  • [Ste05] S. Stevens. Semisimple characters for p-adic classical groups. Duke Math. J., 127(1):123–173, 2005.
  • [Wei82] A. Weil. Adeles and algebraic groups, volume 23 of Prog. in Math. Birkhäuser Boston, Mass., 1982.