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

On Flips of Unitary Buildings I: Classification of Flips

Abstract

We classify flips of buildings arising from non-degenerate unitary spaces of dimension at least 4 over finite fields of odd characteristic in terms of their action on the underlying vector space. We also construct certain geometries related to flips and prove that these geometries are flag transitive.

Rieuwert J. Blok1 and Benjamin Carr1,2
1Department of Mathematics and Statistics
Bowling Green State University
Bowling Green, Ohio 43402 USA
2MCPE Division
Lindenwood University
209 S. Kingshighway
St. Charles, Missouri 63301 USA
bcarr@member.ams.org

Keywords: building, flip, phan involution, incidence geometry


AMS Subject Classification (2010): Primary 51A50; Secondary 51E26


Suggested Running Title: On Flips of Unitary Buildings I

1 Introduction

1.1 History

This paper should be viewed as part of a program described in [2] to prove theorems similar to Phan’s theorem. These so-called “Phan-type” theorems have been studied in a number of papers (e.g. [4], [3], [7]) initially in order to aid the Gorenstein-Lyons-Solomon revision of the proof of the Classification of Finite Simple Groups. Roughtly speaking, these “Phan-type” theorems allow for the recognition of a group based on amalgams of subgroups that are produced by the group acting on a geometry. These results all rely on the fact that if a geometry is simply connected, then a flag transitive automorphism group of the geometry is the universal completion of its amalgam of maximal parabolic subgroups. The reader interested in more detail should consult [2] for an overview.

The strategy to prove further Phan-type theorems is to identify a simply connected flag transitive geometry, and a group acting flag transitively on the geometry. The notion of a flip (or Phan involution to some authors) was introduced in [2] as a means to produce new geometries which are, in many cases, simply connected and flag transitive.

Flips are studied in a more general context in [9] and [8] where their properties are explored, however the authors do not make a closer study of flips of the building under consideration here.

1.2 The Results of This Paper

Throughout this paper, qq denotes an odd prime power, Δ\Delta denotes the building associated to the geometry of totally isotropic subspaces of a 2n2n-dimensional (n2n\geq 2) non-degenerate unitary space (V,β)(V,\beta) over 𝔽=𝔽q2\mathbb{F}=\mathbb{F}_{q^{2}}, and σ\sigma denotes the qqth power map on 𝔽\mathbb{F}.

In this paper we classify flips of Δ\Delta in terms of their action on (V,β)(V,\beta). Since the interest in flips arises because of the possibility of proving further Phan-type theorems our proof is highly geometric, relying on the construction of geometries induced by the flip. Finally, we prove that these geometries are flag transitive and therefore can be used to prove Phan-type theorems when they are simply connected. In [6] we study the topological properties of these geometries and show that in large rank they are simply connected.

The main results of this paper are as follows:

Main Theorem 1: Classification of Flips.

Let φ\varphi be a flip of Δ\Delta. Then φ\varphi is induced by a semilinear transformation ff of the underlying unitary space VV such that exactly one of the following holds:

  1. (i)

    ff is a linear isometry of (V,β)(V,\beta), f2=idf^{2}=\mathrm{id} on VV, and there is a hyperbolic basis {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} for VV such that f(ei)=fif(e_{i})=f_{i} for i=1,,ni=1,\ldots,n;

  2. (ii)

    ff is a linear anti-isometry of (V,β)(V,\beta), f2=idf^{2}=\mathrm{id} on VV, and there is a hyperbolic basis {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} for VV such that f(ei)=αfif(e_{i})=\alpha f_{i} and f(fi)=α1eif(f_{i})=\alpha^{-1}e_{i} for i=1,,ni=1,\ldots,n, where α\alpha is a trace 0 element of 𝔽\mathbb{F};

  3. (iii)

    ff is a σ\sigma-semilinear isometry of (V,β)(V,\beta), f2=idf^{2}=\mathrm{id} on VV, and there is a hyperbolic basis {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} for VV such that f(ei)=fif(e_{i})=f_{i} for i=1,,ni=1,\ldots,n;

  4. (iv)

    ff is a σ\sigma-semilinear isometry of (V,β)(V,\beta), f2=idf^{2}=\mathrm{id} on VV, and there is a hyperbolic basis {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} for VV such that for i=1,,n1i=1,\ldots,n-1, f(ei)=fif(e_{i})=f_{i}, f(fi)=eif(f_{i})=e_{i} and there is a non-square λ𝔽\lambda\in\mathbb{F} with f(en)=λfnf(e_{n})=\lambda f_{n} and f(fn)=σ(λ1)enf(f_{n})=\sigma(\lambda^{-1})e_{n}.

Conversely any semilinear transformation of VV satisfying one of (i)-(iv) induces a flip of Δ\Delta.

So there are up to a unitary base change only four flips of Δ\Delta. Each flip gives rise to non-isomorphic geometries which can be used to prove Phan-type theorems about flag-transitive automorphism groups of the geometries when the geometries are simply connected by appealing to Tits’ Lemma (Corollaire 1 of [14].)

In the body of the paper this theorem is split into four pieces. First we prove in Lemma 3.4 that every flip of Δ\Delta is induced by some linear isometry, linear anti-isometry, or σ\sigma-semilinear isometry of (V,β)(V,\beta). Then, in Lemma 3.9 we prove that a semilinear transformation of VV satisfying any of (i)-(iv) induces a flip of Δ\Delta. We then prove in Main Theorem 1A (Section 4.1) that if the transformation is linear, then (i) or (ii) holds. Finally in Main Theorem 1B (Section 5.2) we prove that if the transformation is σ\sigma-semilinear then either (iii) or (iv) holds.

In addition to classifying the flips of Δ\Delta we prove the following results regarding the geometries Γ(n,q)\Gamma(n,q) and Γ1(n,q)\Gamma_{1}(n,q). The construction of Γ(n,q)\Gamma(n,q) is carried out in Section 3.3 and the construction of Γ1(n,q)\Gamma_{1}(n,q) is carried out in Section 5.1.

Main Theorem 2: Linear Flag Transitivity.

If φ\varphi is a flip of Δ\Delta induced by a linear transformation of (V,β)(V,\beta) then the geometry Γ(n,q)\Gamma(n,q) is flag transitive.

Main Theorem 3: σ\sigma-Semilinear Flag Transitivity.

If φ\varphi is a flip of Δ\Delta induced by a σ\sigma-semilinear transformation of (V,β)(V,\beta) then the geometry Γ1(n,q)\Gamma_{1}(n,q) is flag transitive.

With these results in hand, the last step to establishing new Phan-type theorems is to study the homotopy properties of these geometries. This is done for large rank cases in [6].

1.3 Acknowledgments

The results of this paper are part of the second authors Ph.D. thesis, [5], under the supervision of the first author. We would also like to express our gratitude to Professor Antonio Pasini for a careful proofreading of the paper and his many helpful comments. Finally we would like to thank the anonymous referee for his comments.

2 Definitions

2.1 Incidence Geometry

Definition 2.1.

Let II be a set. A pregeometry over II is a set Γ\Gamma together with a type function t:ΓIt:\Gamma\to I and a symmetric, reflexive incidence relation \sim on Γ\Gamma with the property that for xx, yΓy\in\Gamma, xyx\sim y and t(x)=t(y)t(x)=t(y) implies x=yx=y. The set II is called the type set of the pregeometry. The cardinality of II is called the rank of the pregeometry. The elements of Γ\Gamma are called the objects of the pregeometry.

A pregeometry is often denoted by an ordered quadruple (Γ,I,t,)(\Gamma,I,t,\sim). If the context is unambiguous the pregeometry may be denoted Γ\Gamma.

Definition 2.2.

Let Γ\Gamma be a pregeometry. A flag is a set of pairwise incident elements. The type of a flag ={Fi1,,Fik}\mathcal{F}=\{F_{i_{1}},\ldots,F_{i_{k}}\} is t()={t(Fij)|j=1,,k}t(\mathcal{F})=\{t(F_{i_{j}})|j=1,\ldots,k\}. The cotype of \mathcal{F} is It()I\setminus t(\mathcal{F}). A flag of type II is called a chamber.

A flag FF is maximal if it is not properly contained in any other flag.

Γ\Gamma is transversal if every maximal flag is a chamber. A transversal pregeometry is called a geometry.

Definition 2.3.

Let FF be a flag in a geometry Γ\Gamma. The residue of FF in Γ\Gamma, denoted resΓ(F)\mathrm{res}_{\Gamma}(F), is the set of all elements of ΓF\Gamma\setminus F that are incident to all elements of FF. The residue of a flag is a geometry with type set It(F)I\setminus t(F). The rank of a residue is called the corank of the flag.

Definition 2.4.

An automorphism of a geometry Γ\Gamma is a permutation of its objects that preserves incidence and type. Denote the group of all automorphisms of Γ\Gamma by Aut(Γ)\mathrm{Aut}(\Gamma).

Definition 2.5.

Let Γ\Gamma be a geometry and let GAut(Γ)G\leq\mathrm{Aut}(\Gamma). We say that GG acts flag transitively on Γ\Gamma if, given two flags CC, DD of Γ\Gamma of the same type, there is an element gGg\in G so that g(C)=Dg(C)=D. If Aut(Γ)\mathrm{Aut}(\Gamma) acts flag transitively on Γ\Gamma then Γ\Gamma is called a flag transitive geometry.

2.2 Buildings and Flips

The material in this section follows [1] with the exception of the definition of a flip, which is taken from [2].

Definition 2.6.

Let (W,S)(W,S) be a Coxeter system. A building of type (W,S)(W,S) is a non-empty set 𝒞\mathcal{C} together with a map δ:𝒞×𝒞W\delta:\mathcal{C}\times\mathcal{C}\to W such that for all C,D𝒞C,D\in\mathcal{C} we have:

  1. (i)

    δ(C,D)=1\delta(C,D)=1 if and only if C=DC=D;

  2. (ii)

    If δ(C,D)=w\delta(C,D)=w and C𝒞C^{\prime}\in\mathcal{C} with δ(C,C)=sS\delta(C^{\prime},C)=s\in S then δ(C,D)=sw\delta(C^{\prime},D)=sw or ww. Moreover if l(sw)=l(w)+1l(sw)=l(w)+1 then δ(C,D)=sw\delta(C^{\prime},D)=sw.

  3. (iii)

    If δ(C,D)=w\delta(C,D)=w then for any sSs\in S there is an element C𝒞C^{\prime}\in\mathcal{C} with δ(C,C)=s\delta(C^{\prime},C)=s and δ(C,D)=sw\delta(C^{\prime},D)=sw.

The elements of 𝒞\mathcal{C} are called chambers.

A building of type (W,S)(W,S) is called spherical if (W,S)(W,S) is a spherical Coxeter system.

Let (𝒞,δ)(\mathcal{C},\delta) be a spherical building of type (W,S)(W,S). Two chambers CC and DD are opposite if δ(C,D)=w0\delta(C,D)=w_{0}, where w0w_{0} is the longest word of (W,S)(W,S).

Definition 2.7.

Let (𝒞,δ)(\mathcal{C},\delta), (𝒞,δ)(\mathcal{C}^{\prime},\delta^{\prime}) be buildings of type (W,S)(W,S). An isomorphism between (𝒞,δ)(\mathcal{C},\delta) and (𝒞,δ)(\mathcal{C}^{\prime},\delta^{\prime}) is a bijection ρ:𝒞𝒞\rho:\mathcal{C}\to\mathcal{C}^{\prime} such that for all uu, v𝒞v\in\mathcal{C}, δ(u,v)=δ(ρ(u),ρ(v))\delta(u,v)=\delta^{\prime}(\rho(u),\rho(v)).

An automorphism of (𝒞,δ)(\mathcal{C},\delta) is an automorphism of (𝒞,δ)(\mathcal{C},\delta) with itself.

Remark. What we have called isomorphisms are sometimes called isometries of the building, with the term isomorphism reserved for a larger class of maps. For the building associated to the geometry of totally isotropic subspaces of a non-degenerate unitary space over a finite field the two terms are equivalent. \lozenge

Definition 2.8.

An apartment of a building (𝒞,δ)(\mathcal{C},\delta) of type (W,S)(W,S) is a subset AA of (𝒞,δ)(\mathcal{C},\delta) such that (A,δ|A)(A,\delta|_{A}) is isomorphic to the Coxeter building of type (W,S)(W,S).

Definition 2.9.

Let (𝒞,δ)(\mathcal{C},\delta) be a spherical building of type (W,S)(W,S) and let w0w_{0} be the longest word of (W,S)(W,S). A flip is a map f:𝒞𝒞f:\mathcal{C}\to\mathcal{C} such that for all C,D𝒞C,D\in\mathcal{C}:

  1. (i)

    f2(C)=Cf^{2}(C)=C;

  2. (ii)

    δ(C,D)=w0δ(f(C),f(D))w0\delta(C,D)=w_{0}\delta(f(C),f(D))w_{0};

  3. (iii)

    There exists C𝒞C\in\mathcal{C} such that δ(C,f(C))=w0\delta(C,f(C))=w_{0}.

Note 2.1.

It follows from (ii) that a flip is an isometry of the building if and only if w0w_{0} is central in WW. In particular this holds for the building studied in this paper.

2.3 The Apartments of Δ\Delta

Recall that Δ\Delta denotes the building associated to the geometry of totally isotropic subspaces of the unitary space (V,β)(V,\beta).

We now describe the apartments of Δ\Delta. This description is valid in a wider context, the interested reader can consult Chapter 7 of [12]. For concreteness we assume that VV is a left vector space over 𝔽\mathbb{F}. Because of this convention we have that for all uu, vVv\in V, λ\lambda, μ𝔽\mu\in\mathbb{F}

β(λu,μv)=λβ(u,v)σ(μ)=λσ(μ)β(u,v).\beta(\lambda u,\mu v)=\lambda\beta(u,v)\sigma(\mu)=\lambda\sigma(\mu)\beta(u,v).
Construction 1.

Let U1=e1,,enU_{1}=\langle e_{1},\ldots,e_{n}\rangle be a maximal totally isotropic subspace of VV. It can be shown (see for example Lemma 7.5 of [12]) that there is a totally isotropic subspace U2=f1,,fnU_{2}=\langle f_{1},\ldots,f_{n}\rangle such that (e1,f1),,(en,fn)(e_{1},f_{1}),\ldots,(e_{n},f_{n}) are pairwise orthogonal hyperbolic pairs, i.e. β(ei,fj)=δij\beta(e_{i},f_{j})=\delta_{ij}, where δij\delta_{ij} is the Kronecker δ\delta, and β(ei,ej)=β(fi,fj)=0\beta(e_{i},e_{j})=\beta(f_{i},f_{j})=0 for all ii, j=1,,nj=1,\ldots,n. Recall that {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} forms a hyperbolic basis for VV. The polar frame associated to {U1,U2}\{U_{1},U_{2}\} is

={ei,fi|1in}.\mathcal{F}=\{\langle e_{i}\rangle,\langle f_{i}\rangle|1\leq i\leq n\}.

Notice that it is the subspaces ei\langle e_{i}\rangle, fi\langle f_{i}\rangle that define the polar frame, not the particular vectors eie_{i}, fif_{i}. Hence different hyperbolic bases for VV may give rise to the same polar frame.

The apartment of \mathcal{F} in Δ\Delta consists of all flags FF that are spanned by some subset of {e1,f1,,en,fn}\{e_{1},f_{1},\ldots,e_{n},f_{n}\}. This apartment is denoted Σ()\Sigma(\mathcal{F}).

Every apartment of Δ\Delta is of the form Σ()\Sigma(\mathcal{F}) for some polar frame \mathcal{F}.

Note 2.2.

In what follows, if ={ei,fi|i=1,,n}\mathcal{F}=\{\langle e_{i}\rangle,\langle f_{i}\rangle|i=1,\ldots,n\} is a polar frame we denote the apartment Σ()\Sigma(\mathcal{F}) by

Σ()=Σ(ei,fi|i=1,,n)\Sigma(\mathcal{F})=\Sigma(\langle e_{i}\rangle,\langle f_{i}\rangle|i=1,\ldots,n)

or if the context permits,

Σ()=Σ(ei,fi).\Sigma(\mathcal{F})=\Sigma(e_{i},f_{i}).

This last notation is somewhat of an abuse, since the collection of pairwise orthogonal hyperbolic pairs is not uniquely determined by the polar frame, but if we start with this collection we know the frame, and hence the apartment.

Now that we know what the apartments of Σ(Γ)\Sigma(\Gamma) look like, we can describe when two chambers are opposite. The following appears as Exercise 9.16(ii) of [12].

Theorem 2.1.

Two chambers C=(Ci)i=1hC=(C_{i})_{i=1}^{h} and D=(Di)i=1hD=(D_{i})_{i=1}^{h} in the building of a non-degenerate polar geometry (W,ρ)(W,\rho) of rank h>0h>0 are opposite if and only if for all ii,

CiDi={0}.C_{i}^{\perp}\cap D_{i}=\{0\}.

3 First Results on Flips

3.1 The Unitary Building and its Flips

Let Δ\Delta denote the building associated to the polar geometry of (V,β)(V,\beta). It is shown in Chapter 7 of [13] that Δ\Delta is a building of type CnC_{n}.

Definition 3.1.

A similitude of the polar space (V,β)(V,\beta) is a τ\tau-semilinear transformation (τAut(𝔽)\tau\in\mathrm{Aut}(\mathbb{F})) ff of VV with the property that there exists some a=σ(a)𝔽qa=\sigma(a)\in\mathbb{F}_{q} such that for all uu, vVv\in V, β(f(u),f(v))=aτ(β(u,v))\beta(f(u),f(v))=a\tau(\beta(u,v)). If a=1a=1, ff is an isometry. If a=1a=-1 then ff is called an anti-isometry.

The group of all similitudes of (V,β)(V,\beta) is denoted ΓU(V)\Gamma\mathrm{U}(V). By PΓU(V)\mathrm{P}\Gamma\mathrm{U}(V) we denote the quotient of ΓU(V)\Gamma\mathrm{U}(V) by its center.

Theorem 3.1.

Aut(Δ)PΓU(V)\mathrm{Aut}(\Delta)\cong\mathrm{P}\Gamma\mathrm{U}(V).

Sketch of Proof.

Since the polar geometry (V,β)(V,\beta) is embeddable in a projective geometry and dimV4\dim V\geq 4, the Fundamental Theorem of Projective Geometry applies to ensure that every automorphism of the polar space is induced by a semilinear transformation of VV. It follows that the automorphism group of the polar geometry is isomorphic to PΓU(V)\mathrm{P}\Gamma\mathrm{U}(V). Finally that every automorphism of the building arises from an automorphism of the geometry is shown on Page 264 of [11]. ∎

Note 3.1.

The main interest of Theorem 3.1 is that there is a surjective homomorphism ΓU(V)Aut(Δ)\Gamma\mathrm{U}(V)\twoheadrightarrow\mathrm{Aut}(\Delta) so that in the proof of Lemma 3.4 we can argue that any flip of Δ\Delta is induced by some transformation in ΓU(V)\Gamma\mathrm{U}(V).

Lemma 3.2.

Let φ\varphi be a flip of Δ\Delta. Then φ\varphi is induced by a similitude ff of VV which satisfies f2=λidf^{2}=\lambda\mathrm{id} on VV for some scalar λ\lambda.

Proof.

Recall first that the longest word w0w_{0} of the Weyl group of type CnC_{n} is central. Thus a flip φ\varphi in fact satisfies δ(u,v)=δ(φ(u),φ(v))\delta(u,v)=\delta(\varphi(u),\varphi(v)) and so is an automorphism of Δ\Delta. It follows from Theorem 3.1 that φ\varphi is induced by some semilinear map fΓU(V)f\in\Gamma\mathrm{U}(V).

Since φ2=id\varphi^{2}=\mathrm{id} on Δ\Delta we see that f2f^{2} is in the kernel of the action of ΓU(V)\Gamma\mathrm{U}(V) on Δ\Delta, which is Z(V)ΓU(V)Z(V)\cap\Gamma\mathrm{U}(V), the group of scalar transformations that also lie in ΓU(V)\Gamma\mathrm{U}(V). Thus f2=λidf^{2}=\lambda\mathrm{id} on VV for some λ𝔽\lambda\in\mathbb{F}. ∎

Recall that the norm Nσ:𝔽𝔽qN_{\sigma}:\mathbb{F}\to\mathbb{F}_{q} defined by Nσ(x)=xσ(x)N_{\sigma}(x)=x\sigma(x) is surjective since 𝔽\mathbb{F} is finite.

Lemma 3.3.

Let φ\varphi be induced by a similitude ff of VV. Then either ff is linear or ff is σ\sigma-semilinear. Moreover if β(f(u),f(v))=aτ(β(u,v))\beta(f(u),f(v))=a\tau(\beta(u,v)) and f2=λidf^{2}=\lambda\mathrm{id} then Nσ(λ)=a2N_{\sigma}(\lambda)=a^{2}.

Proof.

Suppose ff is τ\tau-semilinear for τAut(𝔽)\tau\in\mathrm{Aut}(\mathbb{F}). Let η𝔽\eta\in\mathbb{F} and let uVu\in V with u0u\neq 0. Since f2=λidf^{2}=\lambda\mathrm{id} on VV it follows that f2(ηu)=ληuf^{2}(\eta u)=\lambda\eta u. But we can calculate directly that

f2(ηu)=f(τ(η)f(u))=τ2(η)f2(u)=τ2(η)λu.f^{2}(\eta u)=f(\tau(\eta)f(u))=\tau^{2}(\eta)f^{2}(u)=\tau^{2}(\eta)\lambda u.

Thus τ2(η)=η\tau^{2}(\eta)=\eta and so τ2\tau^{2} is the identity of Aut(𝔽)\mathrm{Aut}(\mathbb{F}). Since Aut(𝔽)\mathrm{Aut}(\mathbb{F}) contains a unique involution it follows that either τ=id\tau=\mathrm{id} and ff is linear, or τ=σ\tau=\sigma and ff is σ\sigma-semilinear.

In order to prove the second part of the theorem, notice that

Nσ(λ)β(u,v)=β(f2(u),f2(v))=aτ(β(f(u),f(v))).N_{\sigma}(\lambda)\beta(u,v)=\beta(f^{2}(u),f^{2}(v))=a\tau(\beta(f(u),f(v))).

Since β(f(u),f(v))=aτβ(u,v)\beta(f(u),f(v))=a\tau\beta(u,v) it follows that

aτ(β(f(u),f(v)))=aτ(a)τ2(β(u,v))=a2β(u,v).a\tau(\beta(f(u),f(v)))=a\tau(a)\tau^{2}(\beta(u,v))=a^{2}\beta(u,v).

This string of equalities relies on the fact that either τ=id\tau=\mathrm{id} or τ=σ\tau=\sigma, and in either case τ(a)=a\tau(a)=a.

Putting these two strings of equalities together we see that for all uu, vVv\in V,

Nσ(λ)β(u,v)=a2β(u,v)N_{\sigma}(\lambda)\beta(u,v)=a^{2}\beta(u,v)

and so since β\beta is non-degenerate, Nσ(λ)=a2N_{\sigma}(\lambda)=a^{2}. ∎

Lemma 3.4.

Let φ\varphi be a flip of Δ\Delta. Then one of the following holds:

  1. (i)

    φ\varphi is induced by a linear isometry fU(V)f\in\mathrm{U}(V) satisfying f2=idf^{2}=\mathrm{id} on VV; or

  2. (ii)

    φ\varphi is induced by a linear anti-isometry ff of VV satisfying f2=idf^{2}=\mathrm{id} on VV; or

  3. (iii)

    φ\varphi is induced by a σ\sigma-semilinear isometry fΓU(V)f\in\Gamma\mathrm{U}(V) so that f2=idf^{2}=\mathrm{id} on VV.

Proof.

By Lemma 3.2 φ\varphi is induced by a similitude ff of VV with f2=λidf^{2}=\lambda\mathrm{id} on VV for some scalar λ𝔽\lambda\in\mathbb{F}.

Since φ\varphi maps some chamber of Δ\Delta to an opposite, there is an apartment

Σ=Σ(ei,fi|i=1,,n)\Sigma=\Sigma(e_{i},f_{i}|i=1,\ldots,n)

in which φ\varphi sends the chamber C=(Ci)i=1nC=(C_{i})_{i=1}^{n} defined by Ci=e1,,eiC_{i}=\langle e_{1},\ldots,e_{i}\rangle to its opposite in Σ\Sigma, the chamber D=(Di)i=1nD=(D_{i})_{i=1}^{n} defined by Di=f1,,fiD_{i}=\langle f_{1},\ldots,f_{i}\rangle.

Since CC and DD are opposite, they lie in a unique apartment. It follows that φ\varphi preserves the apartment Σ\Sigma. In particular, since for each i=1,,ni=1,\ldots,n we have ei=CiDi1\langle e_{i}\rangle=C_{i}\cap D_{i-1}^{\perp} and fi=DiCi1\langle f_{i}\rangle=D_{i}\cap C_{i-1}^{\perp} we see that φ\varphi sends each 1-object to its opposite in Σ\Sigma and so for each i=1,,ni=1,\ldots,n there exist scalars λi,μi𝔽\lambda_{i},\mu_{i}\in\mathbb{F} so that

f(ei)\displaystyle f(e_{i}) =\displaystyle= λifi\displaystyle\lambda_{i}f_{i}
f(fi)\displaystyle f(f_{i}) =\displaystyle= μiei.\displaystyle\mu_{i}e_{i}.
  1. (a)

    Suppose ff is linear and for all uu, vVv\in V, β(f(u),f(v))=aβ(u,v)\beta(f(u),f(v))=a\beta(u,v). Since a𝔽qa\in\mathbb{F}_{q} there exists μ𝔽\mu\in\mathbb{F} such that Nσ(μ)=a1N_{\sigma}(\mu)=a^{-1}. Replace ff by μf\mu f and we see that for all uu, vVv\in V,

    β((μf)(u),(μf)(v))=Nσ(μ)β(f(u),f(v))=a1aβ(u,v)=β(u,v).\beta((\mu f)(u),(\mu f)(v))=N_{\sigma}(\mu)\beta(f(u),f(v))=a^{-1}a\beta(u,v)=\beta(u,v).

    Thus μf\mu f is an isometry which also induces φ\varphi.

    Suppose now that we have chosen an isometry ff which induces φ\varphi, and f2=λidf^{2}=\lambda\mathrm{id}. If λ=1\lambda=1 the the conclusion of (i) is satisfied and we’re done. So assume λ1\lambda\neq 1. Notice that we have the following equalities:

    β(u,v)=β(f(u),f(v))\displaystyle\beta(u,v)=\beta(f(u),f(v)) =\displaystyle= β(f2(u),f2(v))=Nσ(λ)β(u,v)\displaystyle\beta(f^{2}(u),f^{2}(v))=N_{\sigma}(\lambda)\beta(u,v) (1)
    σ(λ1)=β(e1,f(e1))\displaystyle\sigma(\lambda_{1})=\beta(e_{1},f(e_{1})) =\displaystyle= β(f(e1),f2(e1))=β(λ1f1,λe1)=λ1σ(λ).\displaystyle\beta(f(e_{1}),f^{2}(e_{1}))=\beta(\lambda_{1}f_{1},\lambda e_{1})=\lambda_{1}\sigma(\lambda). (2)

    It follows from (1) that Nσ(λ)=1N_{\sigma}(\lambda)=1, and from (2) that λ\lambda is a square in 𝔽\mathbb{F}. Choose η𝔽\eta\in\mathbb{F} so that η2=λ1\eta^{2}=\lambda^{-1}. Since NσN_{\sigma} is multiplicative, it follows that Nσ(η)2=Nσ(η2)=Nσ(λ)=1N_{\sigma}(\eta)^{2}=N_{\sigma}(\eta^{2})=N_{\sigma}(\lambda)=1 and so Nσ(η){±1}N_{\sigma}(\eta)\in\{\pm 1\}.

    Let g=ηfg=\eta f. Then g2=idg^{2}=\mathrm{id} on VV, but we have paid a price. We now have that

    β(g(u),g(v))=β(ηf(u),ηf(v))=Nσ(η)β(u,v).\beta(g(u),g(v))=\beta(\eta f(u),\eta f(v))=N_{\sigma}(\eta)\beta(u,v).

    Thus either gg is an isometry of (V,β)(V,\beta) or gg is an anti-isometry of (V,β)(V,\beta). If gg is an isometry the conclusion of (i) is satisfied, and if gg is an anti-isometry the conclusion of (ii) is satisfied.

  2. (b)

    Suppose now that ff is semilinear but not linear. Then by Lemma 3.3 ff is σ\sigma-semilinear. We now show that we can replace ff by a scalar multiple μf\mu f which still induces φ\varphi so that (μf)2=id(\mu f)^{2}=\mathrm{id} on VV. Namely we have

    λei\displaystyle\lambda e_{i} =\displaystyle= f2(ei)=μiσ(λi)ei\displaystyle f^{2}(e_{i})=\mu_{i}\sigma(\lambda_{i})e_{i}
    λfi\displaystyle\lambda f_{i} =\displaystyle= f2(fi)=λiσ(μi)fi\displaystyle f^{2}(f_{i})=\lambda_{i}\sigma(\mu_{i})f_{i}

    and so λ=μiσ(λi)=λiσ(μi)\lambda=\mu_{i}\sigma(\lambda_{i})=\lambda_{i}\sigma(\mu_{i}). Hence λ\lambda lies in 𝔽σ=𝔽q\mathbb{F}^{\sigma}=\mathbb{F}_{q}, the fixed field of σ\sigma. Since 𝔽\mathbb{F} is finite the norm map Nσ:𝔽𝔽qN_{\sigma}:\mathbb{F}\to\mathbb{F}_{q} is surjective. Thus there exists μ𝔽\mu\in\mathbb{F} so that Nσ(μ)=λ1N_{\sigma}(\mu)=\lambda^{-1}. Replacing ff by μf\mu f does not affect φ\varphi, and so we may do this and assume λ=1\lambda=1.

    In order to check that ff can be taken to be an isometry, by Lemma 3.2, since λ=1\lambda=1 also a2=1a^{2}=1. Hence a{±1}a\in\{\pm 1\}. The following calculation shows that a1=1a^{-1}=1 and so a=1a=1 and we are in the situation of (iii):

    1=β(e1,f1)=a1σ1(β(f(e1),f(f1)))=a1σ1(β(λ1f1,μ1e1))=a1λ=a1.1=\beta(e_{1},f_{1})=a^{-1}\sigma^{-1}(\beta(f(e_{1}),f(f_{1})))=a^{-1}\sigma^{-1}(\beta(\lambda_{1}f_{1},\mu_{1}e_{1}))=a^{-1}\lambda=a^{-1}.\qed

It is easy to see that the three cases in Lemma 3.4 are mutually exclusive.

Note 3.2.

It follows immediately from Lemma 3.3 that if ff induces a flip φ\varphi and f2=idf^{2}=\mathrm{id} on VV then either ff is an isometry or ff is an anti-isometry. What is interesting about Lemma 3.4 is that if ff is linear we have to consider both the isometry and anti-isometry possibilities, whereas if ff is σ\sigma-semilinear we can assume it is an isometry.

Definition 3.2.

Let φ\varphi be a flip of Δ\Delta. We say φ\varphi is linear if it is induced by a linear transformation of VV. We say φ\varphi is σ\sigma-semilinear if it is induced by a σ\sigma-semilinear transformation of VV.

Note 3.3.

From now on we identify φ\varphi with a transformation of VV that induces φ\varphi and satisfies the appropriate conclusion of Lemma 3.4.

We now define a new form on VV that will be important in the study of geometries induced by φ\varphi.

Definition 3.3.

Given fΓU(V)f\in\Gamma\mathrm{U}(V), define βf(u,v)=β(u,f(v))\beta_{f}(u,v)=\beta(u,f(v)).

Lemma 3.5.

Let fΓU(V)f\in\Gamma\mathrm{U}(V) be τ\tau-semilinear, and assume f2=idf^{2}=\mathrm{id}. Then βf\beta_{f} is a non-degenerate, reflexive, στ\sigma\tau-sesquilinear form. In particular,

  1. (i)

    if σ=τ\sigma=\tau, then βf\beta_{f} is a non-degenerate bilinear form, and

  2. (ii)

    if ff is linear, then βf\beta_{f} is a non-degenerate σ\sigma-sesquilinear form.

Proof.

That βf\beta_{f} is non-degenerate follows since ff is bijective. Left homogeneity follows since β\beta is left homogeneous and ff acts in the second argument. Reflexivity and both (i) and (ii) follow from direct calculations. ∎

Note 3.4.

If ff induces a flip φ\varphi as in Lemma 3.4 then the choice ff determines βφ\beta_{\varphi} up to multiplication by ±1\pm 1 if ff is linear, and up to multiplication by a scalar of norm 1 if ff is σ\sigma-semilinear. It follows that choosing different representatives for φ\varphi does not affect the results of Lemmas 3.6 and 3.7.

Lemma 3.6.

Let φ\varphi be a σ\sigma-semilinear flip of Δ\Delta. Then βφ\beta_{\varphi} is a non-degenerate, reflexive, symmetric, bilinear form.

Proof.

All except the fact that βφ\beta_{\varphi} is symmetric follows from Lemma 3.5. That βφ\beta_{\varphi} is symmetric follows from an easy calculation. ∎

Lemma 3.7.

Let φ\varphi be a linear flip of Δ\Delta.

  1. (i)

    If φ\varphi is a linear isometry then βφ\beta_{\varphi} is a non-degenerate, reflexive, σ\sigma-hermitian form.

  2. (ii)

    If φ\varphi is a linear anti-isometry then βφ\beta_{\varphi} is a non-degenerate, reflexive, σ\sigma-antihermitian form.

Proof.

All that remains is to show that in (i) the form is hermitian and in (ii) the form is antihermitian. Both follow from easy calculations. ∎

Definition 3.4.

Given a flip φ\varphi, set Qφ(v)=12βφ(v,v)Q_{\varphi}(v)=\frac{1}{2}\beta_{\varphi}(v,v).

Notice that QφQ_{\varphi} is the pseudo-quadratic form that polarizes to βφ\beta_{\varphi}.

3.2 The Chamber System Induced by a Flip

We now define a chamber system left invariant by a flip. We shall use this to classify σ\sigma-semilinear flips, but we will also be interested in these for their automorphism groups.

Definition 3.5.

By Δφ\Delta^{\varphi} we denote the collection of chambers of Δ\Delta sent to an opposite chamber by φ\varphi.

Definition 3.6.

Recall that a pair of vectors uu, vVv\in V are β\beta-orthogonal if β(u,v)=0\beta(u,v)=0, this is denoted uvu\perp v. The vectors are βφ\beta_{\varphi}-orthogonal if βφ(u,v)=0\beta_{\varphi}(u,v)=0, this is denoted uφvu\perp_{\varphi}v. The vectors are biorthogonal if β(u,v)=βφ(u,v)=0\beta(u,v)=\beta_{\varphi}(u,v)=0, this is denoted uvu\perp\hskip-4.26773pt\perp v. If UU is a subspace of VV we use UU^{\perp}, UφU^{\perp_{\varphi}}, and UU^{\perp\hskip-4.26773pt\perp} to refer to the β\beta-orthogonal complement, βφ\beta_{\varphi}-orthogonal complement, and biorthogonal complement respectively.

Recall that a pair of β\beta isotropic vectors uu, vv is called a hyperbolic pair if β(u,v)=1\beta(u,v)=1. We define a pre-hyperbolic pair to be a pair of β\beta isotropic vectors uu, vv with β(u,v)0\beta(u,v)\neq 0. This is not standard, but there are instances where the distinction will be important.

Theorem 3.8.

Let φ\varphi be a flip of the unitary building Δ\Delta.

  1. (1)

    A chamber C=(Ci)i=1nC=(C_{i})_{i=1}^{n} of Δ\Delta lies in Δφ\Delta^{\varphi} if and only if CiC_{i} is non-degenerate with respect to βφ\beta_{\varphi} for all i=1,,ni=1,\ldots,n.

  2. (2)

    If {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} is a β\beta pre-hyperbolic basis for VV with φ(ei)=fi\varphi(e_{i})=f_{i}, then the chambers (Ci)i=1n(C_{i})_{i=1}^{n} and (Di)i=1n(D_{i})_{i=1}^{n} defined by Ci=e1,,eiC_{i}=\langle e_{1},\ldots,e_{i}\rangle and Di=f1,,fiD_{i}=\langle f_{1},\ldots,f_{i}\rangle are opposite in Δ\Delta and so lie in Δφ\Delta^{\varphi}. Conversely if C=(Ci)i=1nC=(C_{i})_{i=1}^{n} is a chamber of Δφ\Delta^{\varphi}, then there is a β\beta pre-hyperbolic basis {ei,fi}\{e_{i},f_{i}\} for VV so that Ci=e1,,eiC_{i}=\langle e_{1},\ldots,e_{i}\rangle, φ(ei)=fi\varphi(e_{i})=f_{i} for all i=1,,ni=1,\ldots,n, and the chamber D=(Di)i=1nD=(D_{i})_{i=1}^{n} defined by Di=f1,,fiD_{i}=\langle f_{1},\ldots,f_{i}\rangle lies in Δφ\Delta^{\varphi} and is opposite to CC.

Proof.
  1. (1)

    By assumption we may view φ\varphi as acting on the vector space VV, and have φ2=id\varphi^{2}=\mathrm{id} on VV. Suppose C=(Ci)i=1nC=(C_{i})_{i=1}^{n} is a chamber of Δφ\Delta^{\varphi}. Then since CC is also a chamber of Δ\Delta, each CiC_{i} is β\beta isotropic.

    Recall from Theorem 2.1 that CC is opposite to φ(C)\varphi(C) in Δ\Delta if and only if for each ii,

    φ(Ci)Ci={0}.\varphi(C_{i})\cap C_{i}^{\perp}=\{0\}.

    Notice that

    φ(Radβφ(Ci))=φ(CiCiφ)=φ(Ci)Ci\varphi(\mathrm{Rad}_{\beta_{\varphi}}(C_{i}))=\varphi(C_{i}\cap C_{i}^{\perp_{\varphi}})=\varphi(C_{i})\cap C_{i}^{\perp}

    where the last equality is justified since φ\varphi is a bijective transformation of VV. Thus the βφ\beta_{\varphi} radical of CiC_{i} is {0}\{0\} if and only if φ(Ci)Ci={0}\varphi(C_{i})\cap C_{i}^{\perp}=\{0\}.

  2. (2)

    The first part follows from (1) by noting that if {ei,fi}\{e_{i},f_{i}\} is a β\beta pre-hyperbolic basis for VV with fi=φ(ei)f_{i}=\varphi(e_{i}) for all ii, then for each ii, {e1,,ei}\{e_{1},\ldots,e_{i}\} and {f1,,fi}\{f_{1},\ldots,f_{i}\} form βφ\beta_{\varphi} orthogonal bases for CiC_{i} and DiD_{i} respectively, and satisfy the hypotheses of (1).

    Conversely suppose C=(Ci)i=1nC=(C_{i})_{i=1}^{n} is a chamber of Δφ\Delta^{\varphi}. Choose e1,,ene_{1},\ldots,e_{n} as follows. Pick e1C1{0}e_{1}\in C_{1}-\{0\}. Then pick eiCiφ(Ci1)e_{i}\in C_{i}\cap\varphi(C_{i-1})^{\perp}. The vectors e1,,ene_{1},\ldots,e_{n} are pairwise biorthogonal. Moreover none can be βφ\beta_{\varphi} isotropic as this would contradict the βφ\beta_{\varphi} non-degeneracy of CiC_{i}. Finally, define fi=φ(ei)f_{i}=\varphi(e_{i}) for i=1,,ni=1,\ldots,n. Then {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} gives the desired basis.∎

Lemma 3.9.

If fΓU(V)f\in\Gamma\mathrm{U}(V) satisfies any of (i)-(iv) in the statement of Main Theorem 1 then ff induces a flip of Δ\Delta.

Proof.

Since fΓU(V)f\in\Gamma\mathrm{U}(V), ff induces an automorphism of Δ\Delta and by assumption ff has order 2. It therefore suffices to show that ff maps some chamber of Δ\Delta to an opposite chamber. Let {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} be a hyperbolic basis for VV as in the hypotheses of Main Theorem 1. Let Ci=e1,,eiC_{i}=\langle e_{1},\ldots,e_{i}\rangle for i=1,,ni=1,\ldots,n. Then φ(Ci)=f1,,fi\varphi(C_{i})=\langle f_{1},\ldots,f_{i}\rangle, and clearly Ciφ(Ci)={0}C_{i}^{\perp}\cap\varphi(C_{i})=\{0\} for all ii. Hence the chamber C=(Ci)i=1nC=(C_{i})_{i=1}^{n} is sent to an opposite chamber by ff, and so ff induces a flip of Δ\Delta. ∎

3.3 The Geometry Induced by a Flip

We now define the geometry corresponding to the chamber system induced by a flip.

Definition 3.7.

Let Γ(n,q)\Gamma(n,q) denote the set of all β\beta-isotropic and βφ\beta_{\varphi} non-degenerate subspaces UU of VV. Let I={1,,n}I=\{1,\ldots,n\} and define τ:Γ(n,q)I\tau:\Gamma(n,q)\to I by τ(U)=dim(U)\tau(U)=\dim(U). Finally, define a relation \sim on Γ(n,q)\Gamma(n,q) by UWU\sim W if UWU\subseteq W or WUW\subseteq U. (Γ(n,q),τ,I,)(\Gamma(n,q),\tau,I,\sim) is the geometry induced by φ\varphi.

Lemma 3.10.

Let UU be a β\beta-isotropic subspace of VV. Then UΓ(n,q)U\in\Gamma(n,q) if and only if Uφ(U)={0}U^{\perp}\cap\varphi(U)=\{0\}.

Proof.

UU is βφ\beta_{\varphi} non-degenerate if and only if Uφ(U)={0}U^{\perp}\cap\varphi(U)=\{0\}. ∎

Note 3.5.

It is clear that (Γ(n,q),I,τ,)(\Gamma(n,q),I,\tau,\sim) is a pregeometry, since Γ(n,q)\Gamma(n,q) is a subset of the set of objects of the full projective geometry of VV, 𝒫(V)\mathcal{P}(V), and we have inherited the type and incidence structure from 𝒫(V)\mathcal{P}(V). We will prove in Theorem 3.28 that Γ(n,q)\Gamma(n,q) is in fact a geometry. In order to achieve this goal we will have to study the properties of the vector space endowed with both forms β\beta and βφ\beta_{\varphi} in more detail.

Definition 3.8.

A point of Γ(n,q)\Gamma(n,q) is an object of Γ(n,q)\Gamma(n,q) of type 1. A line of Γ(n,q)\Gamma(n,q) is an object of Γ(n,q)\Gamma(n,q) of type 2.

Note 3.6.

From now on, we have chosen to identify a point of the geometry, which is really a 1-dimensional subspace of VV, with a non-zero vector in that subspace.

The proofs of Lemmata 3.11 and 3.12 are straightforward.

Lemma 3.11.

Let UU be a subspace of VV. Then U=φ(U)φU^{\perp}=\varphi(U)^{\perp_{\varphi}} and

U,φ(U)=U,φ(U)φ=U.\langle U,\varphi(U)\rangle^{\perp}=\langle U,\varphi(U)\rangle^{\perp_{\varphi}}=U^{\perp\hskip-4.26773pt\perp}.
Lemma 3.12.

Let UU, UΓ(n,q)U^{\prime}\in\Gamma(n,q) with UUU\subset U^{\prime}. Then

U,φ(U)U=φ(U)U=UφU=UU.\langle U,\varphi(U)\rangle^{\perp}\cap U^{\prime}=\varphi(U)^{\perp}\cap U^{\prime}=U^{\perp_{\varphi}}\cap U^{\prime}=U^{\perp\hskip-4.26773pt\perp}\cap U^{\prime}.
Lemma 3.13.

Let UU, UΓ(n,q)U^{\prime}\in\Gamma(n,q) with UUU\subset U^{\prime}. Then

W=U,φ(U)UΓ(n,q).W=\langle U,\varphi(U)\rangle^{\perp}\cap U^{\prime}\in\Gamma(n,q).
Proof.

Since WUW\subset U^{\prime} and UU^{\prime} is β\beta isotropic, also WW is β\beta isotropic. By Lemma 3.12 W=UφUW=U^{\perp_{\varphi}}\cap U^{\prime} and so WW is a βφ\beta_{\varphi} orthogonal complement to UU in UU^{\prime}. Since both UU and UU^{\prime} are βφ\beta_{\varphi} non-degenerate it follows that WW is βφ\beta_{\varphi} non-degenerate. ∎

Lemma 3.14.

Let UΓ(n,q)U\in\Gamma(n,q). Then UU contains a point of Γ(n,q)\Gamma(n,q).

Proof.

If dimU=1\dim U=1 then UU is a point. Suppose dimU>1\dim U>1. Since UΓ(n,q)U\in\Gamma(n,q) it is βφ\beta_{\varphi} non-degenerate and so there exists uu, vUv\in U so that βφ(u,v)0\beta_{\varphi}(u,v)\neq 0. If either Qφ(u)0Q_{\varphi}(u)\neq 0 or Qφ(v)0Q_{\varphi}(v)\neq 0 then uu or vv is a point respectively. Otherwise it is straightforward to check that there exists λ𝔽\lambda\in\mathbb{F} so that u+λvu+\lambda v is a point of Γ(n,q)\Gamma(n,q). ∎

3.4 Further Properties of β\beta and βφ\beta_{\varphi}

In this section we have collected some results concerning the relationship between β\beta and βφ\beta_{\varphi}. These results hold for both linear and σ\sigma-semilinear flips and will be used in showing that Γ(n,q)\Gamma(n,q) is a geometry.

Recall that φ\varphi denotes both a flip and a semilinear transformation of VV that induces the flip and satisfies the appropriate conclusion of Lemma 3.4.

Lemma 3.15.

Let UU be a φ\varphi-invariant subspace of VV. Then

Radβ(U)=Radβφ(U)=φ(Radβ(U)).\mathrm{Rad}_{\beta}(U)=\mathrm{Rad}_{\beta_{\varphi}}(U)=\varphi(\mathrm{Rad}_{\beta}(U)).
Proof.

A vector uu lies in the β\beta radical of UU if and only if β(u,v)=0\beta(u,v)=0 for all vUv\in U. Since UU is φ\varphi-invariant, U=φ(U)U=\varphi(U) and so this is also equivalent to requiring that βφ(u,v)=0\beta_{\varphi}(u,v)=0 for all vUv\in U. Thus a vector lies in the β\beta radical of UU if and only if it lies in the βφ\beta_{\varphi} radical of UU. ∎

Note 3.7.

From now on, when referring to the radical of a φ\varphi-invariant subspace we need not specify to which form we are referring.

Lemma 3.16.

Let UU be a φ\varphi-invariant subspace of VV, and let RR be a φ\varphi-invariant subspace of UU. Then RR has a φ\varphi-invariant complement in UU.

Proof.

This is a special case of Maschke’s Theorem, see for example Theorem 1.9 of [10]. ∎

Combining Lemmata 3.15 and 3.16 we immediately find the following.

Corollary 3.17.

Let UU be a φ\varphi-invariant subspace of VV and let RR be its radical. Then RR has a φ\varphi-invariant complement in UU.

Lemma 3.18.

Let WΓ(n,q)W\in\Gamma(n,q) with dimW=k\dim W=k. Then Wφ(W)={0}W\cap\varphi(W)=\{0\}. Hence W=W,φ(W)W^{\prime}=\langle W,\varphi(W)\rangle is 2k2k-dimensional, φ\varphi-invariant, and non-degenerate.

Proof.

Since WW is β\beta isotropic, WWW\subseteq W^{\perp}. Since WW is βφ\beta_{\varphi} non-degenerate it follows that

Wφ(W)={0}W^{\perp}\cap\varphi(W)=\{0\}

and so also Wφ(W)={0}W\cap\varphi(W)=\{0\}. This shows that dimW,φ(W)=2k\dim\langle W,\varphi(W)\rangle=2k.

That WW^{\prime} is φ\varphi-invariant is clear. To show that WW^{\prime} is non-degenerate, notice first that by Lemma 3.15, Radβ(W)=φ(Radβ(W))\mathrm{Rad}_{\beta}(W^{\prime})=\varphi(\mathrm{Rad}_{\beta}(W^{\prime})). Furthermore, Radβ(W)WW=W\mathrm{Rad}_{\beta}(W^{\prime})\subseteq W^{\perp}\cap W^{\prime}=W. But also φ(Radβ(W))φ(WW)=φ(WW)=φ(W)φ(W)=φ(W)\varphi(\mathrm{Rad}_{\beta}(W^{\prime}))\subseteq\varphi(W^{\perp}\cap W^{\prime})=\varphi(W\cap W^{\prime})=\varphi(W)\cap\varphi(W^{\prime})=\varphi(W). It follows that Radβ(W)Wφ(W)={0}\mathrm{Rad}_{\beta}(W^{\prime})\subseteq W\cap\varphi(W)=\{0\} and so WW^{\prime} is non-degenerate. ∎

Corollary 3.19.

If WΓ(n,q)W\in\Gamma(n,q) with dimW=k\dim W=k then there is a basis {wi}i=1k\{w_{i}\}_{i=1}^{k} for WW of biorthogonal points.

Proof.

We induct on kk. If k=1k=1 the result is trivial. If k>1k>1 then by Lemma 3.14 WW contains a point w1w_{1} of Γ(n,q)\Gamma(n,q). Let W=w1WW^{\prime}=\langle w_{1}\rangle^{\perp\hskip-4.26773pt\perp}\cap W. It follows from Lemma 3.12 that W=w1φWW^{\prime}=\langle w_{1}\rangle^{\perp_{\varphi}}\cap W, and so WW^{\prime} has codimension 1 in WW. By Lemma 3.13 WΓ(n,q)W^{\prime}\in\Gamma(n,q) and so by the inductive hypothesis there exists a collection of biorthogonal points {w2,,wk}\{w_{2},\ldots,w_{k}\} that is a basis for WW^{\prime}. Since Ww1,φ(w1)W^{\prime}\subset\langle w_{1},\varphi(w_{1})\rangle^{\perp} it follows that {w1,,wk}\{w_{1},\ldots,w_{k}\} forms a basis of biorthogonal points for WW. ∎

3.5 Γ(n,q)\Gamma(n,q) is a geometry

In this section we prove that Γ(n,q)\Gamma(n,q) is a geometry. Throughout this section φ\varphi denotes both a flip, and a semilinear transformation of VV that induces the flip and satisfies the appropriate conclusion of Lemma 3.4. Unless otherwise stated these results hold for both linear and σ\sigma-semilinear flips.

Lemma 3.20.

Let UU be a subspace of VV with dimU>n\dim U>n. Then φ\varphi does not act as a scalar on UU.

Proof.

Let MM be an nn-dimensional β\beta isotropic βφ\beta_{\varphi} non-degenerate subspace of VV. Then by Lemma 3.18, Mφ(M)={0}M\cap\varphi(M)=\{0\}. If UU is a subspace of dimension greater than nn and φ\varphi acts as a scalar on UU, then φ\varphi acts as a scalar on MU{0}M\cap U\neq\{0\}. Thus there is a non-zero vector vMUv\in M\cap U with φ(v)=μv\varphi(v)=\mu v for some non-zero μ𝔽\mu\in\mathbb{F}. But then vMφ(M)={0}v\in M\cap\varphi(M)=\{0\}, a contradiction. ∎

Lemma 3.21.

Suppose φ\varphi is a linear flip, and let XX be 2k2k-dimensional, φ\varphi invariant and non-degenerate subspace of VV. Then one of the following three holds:

  1. (i)

    XX contains a point of Γ(n,q)\Gamma(n,q);

  2. (ii)

    φ(x)=x\varphi(x)=x for all xXx\in X;

  3. (iii)

    φ(x)=x\varphi(x)=-x for all xXx\in X.

Proof.

Suppose that XX does not contain any points of Γ(n,q)\Gamma(n,q). We will show that either (ii) or (iii) holds. Since XX is β\beta non-degenerate and even dimensional we can write it as an orthogonal direct sum of β\beta hyperbolic lines,

X=i=1mai,biX=\perp_{i=1}^{m}\langle a_{i},b_{i}\rangle

where each (ai,bi)(a_{i},b_{i}) is a hyperbolic pair.

We proceed now in a series of steps to show that φ\varphi acts on XX as either idX\mathrm{id}_{X} or idX-\mathrm{id}_{X}.

  1. Step 1: If uu, vXv\in X are β\beta-isotropic then β(u,v)=0\beta(u,v)=0 if and only if βφ(u,v)=0\beta_{\varphi}(u,v)=0.

    Proof.

    Notice first that since XX contains no points of Γ(n,q)\Gamma(n,q), Qφ(u)=Qφ(v)=0Q_{\varphi}(u)=Q_{\varphi}(v)=0.

    Suppose β(u,v)=0\beta(u,v)=0 but βφ(u,v)0\beta_{\varphi}(u,v)\neq 0. If φ\varphi is an isometry and λ\lambda is chosen so that Trσ(σ(λ)βφ(u,v))0\mathrm{Tr}_{\sigma}(\sigma(\lambda)\beta_{\varphi}(u,v))\neq 0 then u+λvu+\lambda v is a point of Γ(n,q)\Gamma(n,q). If φ\varphi is an anti-isometry and λ\lambda is chosen so that σ(λ)βφ(u,v)λσ(βφ(u,v))0\sigma(\lambda)\beta_{\varphi}(u,v)-\lambda\sigma(\beta_{\varphi}(u,v))\neq 0 then u+λvu+\lambda v is a point of Γ(n,q)\Gamma(n,q). In either case, such λ\lambda exist and so since by hypothesis XX contains no points of Γ(n,q)\Gamma(n,q) we conclude that if β(u,v)=0\beta(u,v)=0 then βφ(u,v)=0\beta_{\varphi}(u,v)=0.

    Conversely if βφ(u,v)=0\beta_{\varphi}(u,v)=0 but β(u,v)0\beta(u,v)\neq 0 then β(u,φ(v))=0\beta(u,\varphi(v))=0 while βφ(u,φ(v))0\beta_{\varphi}(u,\varphi(v))\neq 0, which we have already shown cannot happen.

    Thus β(u,v)=0\beta(u,v)=0 if and only if βφ(u,v)=0\beta_{\varphi}(u,v)=0 for all β\beta-isotropic uu, vXv\in X. ∎

  2. Step 2: For all i=1,mi=1,\ldots m, φ(ai)ai\varphi(a_{i})\in\langle a_{i}\rangle and φ(bi)bi\varphi(b_{i})\in\langle b_{i}\rangle.

    Proof.

    We perform the calculation only for a1a_{1}, the others are similar. Suppose

    φ(a1)=i=1m(xiai+yibi)\varphi(a_{1})=\sum_{i=1}^{m}(x_{i}a_{i}+y_{i}b_{i})

    for some scalars xix_{i}, yi𝔽y_{i}\in\mathbb{F}, i=1,,mi=1,\ldots,m.

    Since β(bi,a1)=0\beta(b_{i},a_{1})=0 for all i1i\neq 1, also βφ(bi,a1)=0\beta_{\varphi}(b_{i},a_{1})=0 for all i1i\neq 1. But we can calculate that βφ(bi,a1)=σ(xi)\beta_{\varphi}(b_{i},a_{1})=\sigma(x_{i}), and so xi=0x_{i}=0 if i1i\neq 1.

    Similarly for all i1i\neq 1, β(ai,a1)=0\beta(a_{i},a_{1})=0 and so also βφ(ai,a1)=0\beta_{\varphi}(a_{i},a_{1})=0, but βφ(ai,a1)=σ(yi)\beta_{\varphi}(a_{i},a_{1})=\sigma(y_{i}) and so yi=0y_{i}=0.

    Hence φ(a1)=x1a1\varphi(a_{1})=x_{1}a_{1}. ∎

  3. Step 3: For all ii, φ(ai)=ai\varphi(a_{i})=a_{i} or φ(ai)=ai\varphi(a_{i})=-a_{i}. Similarly φ(bi)=bi\varphi(b_{i})=b_{i} or φ(bi)=bi\varphi(b_{i})=-b_{i}.

    Proof.

    We prove the result for aia_{i}, the result for bib_{i} is proved similarly. Since φ2=id\varphi^{2}=\mathrm{id} on VV, φ2(ai)=xi2ai=ai\varphi^{2}(a_{i})=x_{i}^{2}a_{i}=a_{i} and so xi2=1x_{i}^{2}=1. Hence xi{±1}x_{i}\in\{\pm 1\}. ∎

  4. Step 4: φ(ai)=ai\varphi(a_{i})=-a_{i} if and only if φ(bi)=bi\varphi(b_{i})=-b_{i}.

    Proof.

    Assume first that φ\varphi is an isometry of (V,β)(V,\beta) and that φ(ai)=ai\varphi(a_{i})=-a_{i}. Then

    1=β(ai,bi)=β(φ(ai),φ(bi))=β(ai,φ(bi))1=\beta(a_{i},b_{i})=\beta(\varphi(a_{i}),\varphi(b_{i}))=-\beta(a_{i},\varphi(b_{i}))

    which forces φ(bi)=bi\varphi(b_{i})=-b_{i}. Similarly if φ(bi)=bi\varphi(b_{i})=-b_{i} then φ(ai)=ai\varphi(a_{i})=-a_{i}.

    Assume next that φ\varphi is an anti-isometry of (V,β)(V,\beta) and that φ(ai)=ai\varphi(a_{i})=-a_{i} while φ(bi)=bi\varphi(b_{i})=b_{i}. Consider the vector x=ai+λbix=a_{i}+\lambda b_{i} where λ\lambda is any non-zero element of trace 0 in 𝔽\mathbb{F}. An easy calculation shows that β(x,x)=0\beta(x,x)=0 while βφ(x,x)=2λ0\beta_{\varphi}(x,x)=-2\lambda\neq 0. Thus xx is a point of Γ(n,q)\Gamma(n,q) which lies in XX, contradicting the assumption that XX contains no points of Γ(n,q)\Gamma(n,q). ∎

  5. Step 5: If φ(a1)=a1\varphi(a_{1})=a_{1} then φ(ai)=ai\varphi(a_{i})=a_{i} for all ii and if φ(a1)=a1\varphi(a_{1})=-a_{1} then φ(ai)=ai\varphi(a_{i})=-a_{i} for all ii.

    Proof.

    Suppose that φ(a1)=a1\varphi(a_{1})=a_{1} but φ(ai)=ai\varphi(a_{i})=-a_{i} for some ii. Then also φ(b1)=b1\varphi(b_{1})=b_{1} and φ(bi)=bi\varphi(b_{i})=-b_{i}. Let x=a1+b1+aibix=a_{1}+b_{1}+a_{i}-b_{i}. Then two easy calculations show that xx is a point of Γ(n,q)\Gamma(n,q). Since by assumption XX contains no points of Γ(n,q)\Gamma(n,q) we conclude that if φ(a1)=a1\varphi(a_{1})=a_{1} then φ(ai)=ai\varphi(a_{i})=a_{i} for all ii. Similarly if φ(a1)=a1\varphi(a_{1})=-a_{1} then φ(ai)=ai\varphi(a_{i})=-a_{i} for all ii. ∎

Thus for all xXx\in X, either φ(x)=x\varphi(x)=x or φ(x)=x\varphi(x)=-x. ∎

The situation is even better for a σ\sigma-semilinear flip:

Lemma 3.22.

Suppose φ\varphi is σ\sigma-semilinear and let UU be a 2k2k-dimensional (k1k\geq 1) subspace of VV that is β\beta non-degenerate. Then either UU is βφ\beta_{\varphi} totally singular or UU contains a point of Γ(n,q)\Gamma(n,q).

Proof.

Assume UU is not βφ\beta_{\varphi} totally singular. Since UU is β\beta non-degenerate we can write U=i=1kai,biU=\perp_{i=1}^{k}\langle a_{i},b_{i}\rangle where each (ai,bi)(a_{i},b_{i}) is a hyperbolic pair. If any aia_{i} or bjb_{j} is a point of Γ(n,q)\Gamma(n,q) then it is the desired point. So we may assume that for all ii, Qφ(ai)=Qφ(bi)=0Q_{\varphi}(a_{i})=Q_{\varphi}(b_{i})=0.

Since UU is not βφ\beta_{\varphi} totally singular we must have one of the following.

  1. (i)

    There is some ii so that βφ(ai,bi)0\beta_{\varphi}(a_{i},b_{i})\neq 0. Then for any non-zero λ\lambda of trace 0, ai+λbia_{i}+\lambda b_{i} is a point of Γ(n,q)\Gamma(n,q).

  2. (ii)

    There are ii, jj so that βφ(ai,aj)0\beta_{\varphi}(a_{i},a_{j})\neq 0. Then ai+aja_{i}+a_{j} is a point of Γ(n,q)\Gamma(n,q).

  3. (iii)

    There are ii, jj so that βφ(ai,bj)0\beta_{\varphi}(a_{i},b_{j})\neq 0. Then ai+bja_{i}+b_{j} is a point of Γ(n,q)\Gamma(n,q).

  4. (iv)

    There are ii, jj so that βφ(bi,bj)0\beta_{\varphi}(b_{i},b_{j})\neq 0, Then bi+bjb_{i}+b_{j} is a point of Γ(n,q)\Gamma(n,q).∎

Corollary 3.23.

Suppose φ\varphi is σ\sigma-semilinear and UU is a 2k2k-dimensional (k1k\geq 1) subspace of VV that is φ\varphi-invariant and non-degenerate. Then UU contains a point of Γ(n,q)\Gamma(n,q).

Theorem 3.24.

Let UΓ(n,q)U\in\Gamma(n,q). If dimU<n\dim U<n then the space X=U,φ(U)X=\langle U,\varphi(U)\rangle^{\perp} contains a point of Γ(n,q)\Gamma(n,q).

Proof.

Notice first that since X=U,φ(U)X^{\perp}=\langle U,\varphi(U)\rangle is non-degenerate by Lemma 3.18, and VV is non-degenerate by hypothesis, also XX is non-degenerate.

If φ\varphi is σ\sigma-semilinear the result now follows immediately from Corollary 3.23.

Now suppose that φ\varphi is linear. We proceed by contradiction. Suppose XX does not contain a point of Γ(n,q)\Gamma(n,q). Then by Lemma 3.21, φ\varphi acts either as idX\mathrm{id}_{X} or idX-\mathrm{id}_{X} on XX. Let k=dimUk=\dim U. Choose a basis {a1,,a2(nk)}\{a_{1},\ldots,a_{2(n-k)}\} for XX.

Let {u1,,uk}\{u_{1},\ldots,u_{k}\} be a basis of biorthogonal points for UU. Recall that such a basis exists by Corollary 3.19. Then

{u1,,uk,φ(u1),,φ(uk)}\{u_{1},\ldots,u_{k},\varphi(u_{1}),\ldots,\varphi(u_{k})\}

forms a basis for U,φ(U)\langle U,\varphi(U)\rangle. We define a new basis for U,φ(U)\langle U,\varphi(U)\rangle by:

{ui+φ(ui),uiφ(ui)|i=1,,k}.\{u_{i}+\varphi(u_{i}),u_{i}-\varphi(u_{i})|i=1,\ldots,k\}.

If φ\varphi acts on XX as idX\mathrm{id}_{X}, define a subspace AA of VV by

A=u1+φ(u1),,uk+φ(uk),a1,,a2(nk).A=\langle u_{1}+\varphi(u_{1}),\ldots,u_{k}+\varphi(u_{k}),a_{1},\ldots,a_{2(n-k)}\rangle.

Then AA is a 2nk>n2n-k>n dimensional subspace of VV on which φ\varphi acts as multiplication by 1, contradicting Lemma 3.20.

Thus by Lemma 3.21, φ\varphi must act on XX as idX-\mathrm{id}_{X}. In this case we define a subspace BB of VV by

B=u1φ(u1),,ukφ(uk),a1,,a2(nk).B=\langle u_{1}-\varphi(u_{1}),\ldots,u_{k}-\varphi(u_{k}),a_{1},\ldots,a_{2(n-k)}\rangle.

Then BB is a 2nk>n2n-k>n dimensional subspace of VV on which φ\varphi acts as multiplication by 1-1, contradicting Lemma 3.20. Hence XX must contain a point of Γ(n,q)\Gamma(n,q). ∎

Corollary 3.25.

If UU is a maximal object of Γ(n,q)\Gamma(n,q) then dimU=n\dim U=n.

Proof.

We proceed by contraposition. If UΓ(n,q)U\in\Gamma(n,q) with dimU<n\dim U<n then by Theorem 3.24 there is a point of Γ(n,q)\Gamma(n,q), uU,φ(U)u\in\langle U,\varphi(U)\rangle^{\perp}. It is easy to see that U,uΓ(n,q)\langle U,u\rangle\in\Gamma(n,q) and so UU is not maximal. ∎

Definition 3.9.

Given an object UΓ(n,q)U\in\Gamma(n,q) and a subspace XX of VV, define rU(X)=XUr_{U}(X)=X\cap U^{\perp\hskip-4.26773pt\perp}.

Lemma 3.26.

Let UU be an mm-object of Γ(n,q)\Gamma(n,q) with m<nm<n and let W=UW=U^{\perp\hskip-4.26773pt\perp}. Then φ|W\varphi|_{W} is a flip of the building of totally isotropic subspaces of (W,β|W)(W,\beta|_{W}).

Proof.

Let MM be a maximal object of Γ(n,q)\Gamma(n,q) containing UU. Let M=WMM^{\prime}=W\cap M. By Corollary 3.19 MM^{\prime} has a basis {m1,,mnm}\{m_{1},\ldots,m_{n-m}\} of biorthogonal points of Γ(n,q)\Gamma(n,q). It is easy to see that {mi,φ(mi)}i=1nm\{m_{i},\varphi(m_{i})\}_{i=1}^{n-m} forms a basis for WW.

Finally, we see that the if for i=1,,nmi=1,\ldots,n-m if we define Di=m1,,miD_{i}=\langle m_{1},\ldots,m_{i}\rangle then D=(Di)i=1nmD=(D_{i})_{i=1}^{n-m} is a chamber of the building of totally isotropic subspaces of (W,β|W)(W,\beta|_{W}) and by Theorem 2.1 we see that DD is sent to its opposite by φ|W\varphi|_{W}. Thus φW\varphi_{W} is a flip of (W,β|W)(W,\beta|_{W}). ∎

We have ignored a subtle point: since there is more than one type of flip, which sort of flip is φ|W\varphi|W? Once we finish proving Main Theorem 1 it will be easy to see that if φ\varphi satisfies (i) or (ii) of Main Theorem 1, then so does φ|W\varphi|_{W}. If φ\varphi satisfies (iii) (resp. (iv)) of Main Theorem 1 and the determinant of the βφ\beta_{\varphi} Gram matrix of UU is a square in 𝔽\mathbb{F} then φ|W\varphi|_{W} also satisfies (iii) (resp. (iv)). If φ\varphi satisfies (iii) (resp. (iv)) and the determinant of the βφ\beta_{\varphi} Gram matrix of UU is a non-square in 𝔽\mathbb{F} then φ|W\varphi|_{W} satisfies (iv) (resp. (iii)).

Corollary 3.27.

If uu is a point of Γ(n,q)\Gamma(n,q) then rur_{u} induces an isomorphism of geometries resΓ(n,q)(u)Γ(n1,q)\mathrm{res}_{\Gamma(n,q)}(u)\to\Gamma(n-1,q).

Proof.

Notice that the objects in the residue of uu correspond to β\beta isotropic βφ\beta_{\varphi} non-degenerate subspaces of W=u,φ(u)W=\langle u,\varphi(u)\rangle^{\perp}, and this correspondence preserves incidence. By Lemma 3.26 φ|W\varphi|_{W} is a flip of (W,β|W)(W,\beta|_{W}) and it is clear from the construction that the geometry induced on WW by φ|W\varphi|_{W} and the geometry on WW induced by φ\varphi agree. Hence resΓ(n,q)(u)Γ(n1,q)\mathrm{res}_{\Gamma(n,q)}(u)\cong\Gamma(n-1,q) and the isomorphism is induced by rur_{u}. ∎

Theorem 3.28.

Γ(n,q)\Gamma(n,q) is a geometry with type and incidence as defined in Definition 3.7.

Proof.

We induct on nn. If n=1n=1 then the result is trivial. Suppose n>1n>1 and let \mathcal{F} be a flag of Γ(n,q)\Gamma(n,q). By Lemma 3.14 we can assume that \mathcal{F} contains a point uu of Γ(n,q)\Gamma(n,q). By Corollary 3.27 the residue of uu is isomorphic to Γ(n1,q)\Gamma(n-1,q), which by the inductive hypothesis is a geometry. Thus ru()r_{u}(\mathcal{F}) is a chamber in Γ(n1,q)\Gamma(n-1,q). It follows easily that \mathcal{F} is a chamber of Γ(n,q)\Gamma(n,q). ∎

With Theorem 3.28 in hand, we can also prove the following:

Lemma 3.29.

Let WW be an object of Γ(n,q)\Gamma(n,q) and let MM be an nn-object of Γ(n,q)\Gamma(n,q) that contains WW. Then any βφ\beta_{\varphi} orthogonal basis for WW extends to a βφ\beta_{\varphi} orthogonal basis for MM. Furthermore if {wi}i=1n\{w_{i}\}_{i=1}^{n} is any βφ\beta_{\varphi} orthogonal basis for MM then {wi,φ(wi)}i=1n\{w_{i},\varphi(w_{i})\}_{i=1}^{n} forms a βφ\beta_{\varphi} orthogonal basis for VV.

Proof.

Notice first that since Γ(n,q)\Gamma(n,q) is a geometry, WW is contained in an nn-dimensional object MM of Γ(n,q)\Gamma(n,q). Let d=dim(W)d=\dim(W) and let {w1,,wm}\{w_{1},\ldots,w_{m}\} be a βφ\beta_{\varphi} orthogonal basis for WW. Then {w1,,wd}\{w_{1},\ldots,w_{d}\} is a basis of biorthogonal points for WW. Let {wd+1,,wn}\{w_{d+1},\ldots,w_{n}\} be a basis of biorthogonal points for W,φ(W)M\langle W,\varphi(W)\rangle^{\perp}\cap M, such a basis exists by combining Lemma 3.13 and Corollary 3.19. Then {w1,,wn}\{w_{1},\ldots,w_{n}\} forms a basis of biorthogonal points for MM, which is in particular a βφ\beta_{\varphi} orthogonal basis for MM.

That {wi,φ(wi)}i=1n\{w_{i},\varphi(w_{i})\}_{i=1}^{n} forms a βφ\beta_{\varphi} orthogonal basis for VV follows since dimM=n\dim M=n and MΓ(n,q)M\in\Gamma(n,q). ∎

Corollary 3.30.

If WW is a kk-object of Γ(n,q)\Gamma(n,q) with βφ\beta_{\varphi} orthogonal basis {wi}i=1k\{w_{i}\}_{i=1}^{k} then there is a basis for VV of β\beta pre-hyperbolic pairs {ei,φ(ei)}i=1n\{e_{i},\varphi(e_{i})\}_{i=1}^{n} with ei=wie_{i}=w_{i} for i=1,,ki=1,\ldots,k.

Proof.

This follows immediately from Lemma 3.29, once one notices that if {e1,,ek}\{e_{1},\ldots,e_{k}\} is a βφ\beta_{\varphi} orthogonal basis for WW then for each i=1,,ki=1,\ldots,k, β(ei,φ(ei))0\beta(e_{i},\varphi(e_{i}))\neq 0 since βφ(ei,ei)0\beta_{\varphi}(e_{i},e_{i})\neq 0. ∎

4 Linear Flips

Throughout this section, φ\varphi denotes a linear flip of Δ\Delta. Recall that we have identified φ\varphi with a linear transformation of VV that induces φ\varphi and satisfies the appropriate conclusion of Lemma 3.4.

4.1 Classification of Linear Flips of the Unitary Building

Main Theorem 1A: Classification of Linear Flips.

Let φ\varphi be a linear flip of Δ\Delta

  1. (i)

    If φ\varphi is induced by an isometry of (V,β)(V,\beta) then there is a basis for VV, {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} of β\beta hyperbolic pairs so that φ(ei)=fi\varphi(e_{i})=f_{i} and φ(fi)=ei\varphi(f_{i})=e_{i} for all i=1,,ni=1,\ldots,n.

  2. (ii)

    If φ\varphi is induced by an anti-isometry of (V,β)(V,\beta) then there is a basis for VV, {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} of β\beta hyperbolic pairs so that φ(ei)=αfi\varphi(e_{i})=\alpha f_{i} and φ(fi)=α1ei\varphi(f_{i})=\alpha^{-1}e_{i} for all i=1,,ni=1,\ldots,n where α\alpha is a trace 0 element of 𝔽\mathbb{F}.

Conversely any linear transformation of VV which satisfies (i) or (ii) induces a flip of Δ\Delta.

Proof.

From the proof of Lemma 3.4 it follows that there is a basis of orthogonal β\beta-hyperbolic pairs {hi,gi}i=1n\{h_{i},g_{i}\}_{i=1}^{n} so that

φ(hi)\displaystyle\varphi(h_{i}) =\displaystyle= λigi, and\displaystyle\lambda_{i}g_{i},\mbox{ and}
φ(gi)\displaystyle\varphi(g_{i}) =\displaystyle= λi1hi\displaystyle\lambda_{i}^{-1}h_{i}

for some λi𝔽\lambda_{i}\in\mathbb{F}.

  1. (i)

    Suppose that φ\varphi is induced by an isometry of (V,β)(V,\beta). Since βφ\beta_{\varphi} is σ\sigma-hermitian it follows that for all i=1,,ni=1,\ldots,n, σ(λi)=βφ(hi,hi)𝔽q\sigma(\lambda_{i})=\beta_{\varphi}(h_{i},h_{i})\in\mathbb{F}_{q} and so in fact σ(λi)=λi\sigma(\lambda_{i})=\lambda_{i}.

    For i=1,,ni=1,\ldots,n let gi=λigig_{i}^{\prime}=\lambda_{i}g_{i}. Choose γi𝔽\gamma_{i}\in\mathbb{F} so that Nσ(γi)=λi1N_{\sigma}(\gamma_{i})=\lambda_{i}^{-1}. Define

    ei\displaystyle e_{i} =\displaystyle= γihi, and\displaystyle\gamma_{i}h_{i},\mbox{ and }
    fi\displaystyle f_{i} =\displaystyle= γigi.\displaystyle\gamma_{i}g_{i}^{\prime}.

    We now calculate to show that {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} is a basis of β\beta hyperbolic pairs with φ(ei)=fi\varphi(e_{i})=f_{i} and φ(fi)=ei\varphi(f_{i})=e_{i}.

    β(ei,ej)\displaystyle\beta(e_{i},e_{j}) =\displaystyle= β(γihi,γjhj)=0\displaystyle\beta(\gamma_{i}h_{i},\gamma_{j}h_{j})=0
    β(fi,fj)\displaystyle\beta(f_{i},f_{j}) =\displaystyle= β(γigi,γigj)=β(γiλigi,γjλjgj)=0\displaystyle\beta(\gamma_{i}g_{i}^{\prime},\gamma_{i}g_{j}^{\prime})=\beta(\gamma_{i}\lambda_{i}g_{i},\gamma_{j}\lambda_{j}g_{j})=0
    β(ei,fi)\displaystyle\beta(e_{i},f_{i}) =\displaystyle= β(γihi,γiλigi)=Nσ(γi)λi=1\displaystyle\beta(\gamma_{i}h_{i},\gamma_{i}\lambda_{i}g_{i})=N_{\sigma}(\gamma_{i})\lambda_{i}=1
    β(ei,fj)\displaystyle\beta(e_{i},f_{j}) =\displaystyle= β(γihi,γjλjgj)=0 if ij.\displaystyle\beta(\gamma_{i}h_{i},\gamma_{j}\lambda_{j}g_{j})=0\mbox{ if }i\neq j.

    Thus {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} forms a β\beta hyperbolic basis for VV, and finally

    φ(ei)=φ(γihi)\displaystyle\varphi(e_{i})=\varphi(\gamma_{i}h_{i}) =\displaystyle= γiφ(hi)=γiλigi=γigi=fi\displaystyle\gamma_{i}\varphi(h_{i})=\gamma_{i}\lambda_{i}g_{i}=\gamma_{i}g_{i}^{\prime}=f_{i}
    φ(fi)=φ(γigi)\displaystyle\varphi(f_{i})=\varphi(\gamma_{i}g_{i}^{\prime}) =\displaystyle= γiφ(gi)=γiλiφ(gi)=γihi=ei.\displaystyle\gamma_{i}\varphi(g_{i}^{\prime})=\gamma_{i}\lambda_{i}\varphi(g_{i})=\gamma_{i}h_{i}=e_{i}.
  2. (ii)

    Suppose now that φ\varphi is induced by an anti-isometry of (V,β)(V,\beta). Since βφ\beta_{\varphi} is σ\sigma-antihermitian it follows that for all ii, σ(λi)=βφ(hi,hi)=λi\sigma(\lambda_{i})=\beta_{\varphi}(h_{i},h_{i})=-\lambda_{i}, and so λi\lambda_{i} is of trace 0. Let α\alpha be any non-zero element of trace 0 in 𝔽\mathbb{F}. For each i=1,,ni=1,\ldots,n choose ai𝔽qa_{i}\in\mathbb{F}_{q} so that aiλi=αa_{i}\lambda_{i}=\alpha. Let γi𝔽\gamma_{i}\in\mathbb{F} be chosen so that Nσ(γi)=aiN_{\sigma}(\gamma_{i})=a_{i}. Set ei=γihie_{i}=\gamma_{i}h_{i} and fi=α1γiλigif_{i}=\alpha^{-1}\gamma_{i}\lambda_{i}g_{i}. Direct calculation shows that {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} is a β\beta hyperbolic basis with the property that φ(ei)=αfi\varphi(e_{i})=\alpha f_{i} and φ(fi)=α1ei\varphi(f_{i})=\alpha^{-1}e_{i}.

The converse follows from Lemma 3.9. ∎

It is now clear that the geometry Γ(n,q)\Gamma(n,q) depends on whether φ\varphi is an isometry or an anti-isometry. With the basis found in Theorem Main Theorem 1A: Classification of Linear Flips we can see that when n=1n=1 the number of points in the geometry depends on whether the flip is an isometry or an anti-isometry, implying that in larger rank the geometries are also not isomorphic.

4.2 A Flag Transitive Automorphism Group of Γ(n,q)\Gamma(n,q)

We are interested in finding a group that acts in a natural way on Γ(n,q)\Gamma(n,q). The obvious choice for this group is the group of linear transformations of VV that preserve both the forms β\beta and βφ\beta_{\varphi}.

Definition 4.1.

Let U2n(q2)φ={fU2n(q2)|βφ(u,v)=βφ(f(u),f(v)) for all u,vV}\mathrm{U}_{2n}(q^{2})^{\varphi}=\{f\in\mathrm{U}_{2n}(q^{2})|\beta_{\varphi}(u,v)=\beta_{\varphi}(f(u),f(v))\mbox{ for all }u,v\in V\}.

In this section we will prove three results regarding U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi}. First we will prove that it is precisely the centralizer in U2n(q2)\mathrm{U}_{2n}(q^{2}) of φ\varphi. We will then prove that it acts flag transitively on Γ(n,q)\Gamma(n,q). We conclude this section by proving that if φ\varphi is induced by an isometry of (V,β)(V,\beta) then U2n(q2)φUn(q2)×Un(q2)\mathrm{U}_{2n}(q^{2})^{\varphi}\cong\mathrm{U}_{n}(q^{2})\times\mathrm{U}_{n}(q^{2}) and if φ\varphi is induced by an anti-isometry then U2n(q2)φGLn(q2)\mathrm{U}_{2n}(q^{2})^{\varphi}\cong\mathrm{GL}_{n}(q^{2}).

Lemma 4.1.

U2n(q2)φ=CΓU2n(q2)(φ)U2n(q2)\mathrm{U}_{2n}(q^{2})^{\varphi}=C_{\Gamma\mathrm{U}_{2n}(q^{2})}(\varphi)\cap\mathrm{U}_{2n}(q^{2}).

Proof.

Let fU2n(q2)φf\in\mathrm{U}_{2n}(q^{2})^{\varphi} and vVv\in V. To check that φ\varphi commutes with ff we will show that for all wVw\in V, β(w,f(φ(v)))=β(w,φ(f(v)))\beta(w,f(\varphi(v)))=\beta(w,\varphi(f(v))). Since β\beta is non-degenerate this will force f(φ(v))=φ(f(v))f(\varphi(v))=\varphi(f(v)).

Let wVw\in V. Choose xVx\in V so that f(x)=wf(x)=w. Then

βφ(x,v)=β(x,φ(v))=β(f(x),f(φ(v)))\displaystyle\beta_{\varphi}(x,v)=\beta(x,\varphi(v))=\beta(f(x),f(\varphi(v))) =\displaystyle= β(w,f(φ(v))) and\displaystyle\beta(w,f(\varphi(v)))\mbox{ and }
βφ(x,v)=βφ(f(x),f(v))=β(f(x),φ(f(v)))\displaystyle\beta_{\varphi}(x,v)=\beta_{\varphi}(f(x),f(v))=\beta(f(x),\varphi(f(v))) =\displaystyle= β(w,φ(f(v))).\displaystyle\beta(w,\varphi(f(v))).

Conversely if fU2n(q2)f\in\mathrm{U}_{2n}(q^{2}) commutes with φ\varphi then for all uu, vVv\in V we have

βφ(f(u),f(v))=β(f(u),φ(f(v)))=β(f(u),f(φ(v)))=β(u,φ(v))=βφ(u,v).\beta_{\varphi}(f(u),f(v))=\beta(f(u),\varphi(f(v)))=\beta(f(u),f(\varphi(v)))=\beta(u,\varphi(v))=\beta_{\varphi}(u,v).\qed

We now turn to the problem of showing that U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} acts flag transitively on Γ(n,q)\Gamma(n,q). Before we can prove that U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} acts flag transitively on Γ(n,q)\Gamma(n,q) we require one more lemma.

Lemma 4.2.

Let C=(Ci)i=1nC=(C_{i})_{i=1}^{n} be a chamber of Γ(n,q)\Gamma(n,q).

  1. (a)

    If φ\varphi is induced by an isometry of (V,β)(V,\beta) as in Lemma 3.4 then there is a basis {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} for VV with the following properties:

    1. (i)

      {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} is hyperbolic with respect to β\beta;

    2. (ii)

      for all i=1,,ni=1,\ldots,n, φ(ei)=fi\varphi(e_{i})=f_{i} and φ(fi)=ei\varphi(f_{i})=e_{i}; and

    3. (iii)

      for all i=1,,ni=1,\ldots,n, Ci=e1,,eiC_{i}=\langle e_{1},\ldots,e_{i}\rangle.

  2. (b)

    If φ\varphi is induced by an anti-isometry of (V,β)(V,\beta) as in Lemma 3.4 then there is a basis {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} for VV with properties (i) and (iii), and

    1. (ii’)

      for all i=1,,ni=1,\ldots,n, φ(ei)=αfi\varphi(e_{i})=\alpha f_{i} and φ(fi)=α1ei\varphi(f_{i})=\alpha^{-1}e_{i} where α\alpha is any non-zero element of trace 0 in 𝔽\mathbb{F}.

Proof.
  1. (a)

    Let e1e_{1} be a non-zero vector in C1C_{1}. Then after scaling as in the proof of Main Theorem 1A(i) we may assume that (e1,φ(e1))(e_{1},\varphi(e_{1})) is a hyperbolic pair. Since C1C2C_{1}^{\perp\hskip-4.26773pt\perp}\cap C_{2} is an element of Γ(n,q)\Gamma(n,q) by Lemma 3.13 we can choose e2C1C2e_{2}\in C_{1}^{\perp\hskip-4.26773pt\perp}\cap C_{2} so that after scaling, (e2,φ(e2))(e_{2},\varphi(e_{2})) is a hyperbolic pair. Repeating this procedure we produce the desired basis.

  2. (b)

    This is proved in the same fashion of (a), with the scaling as in the proof of Main Theorem 1A(ii) replacing the scaling in Main Theorem 1A(i).∎

Main Theorem 2: Linear Flag Transitivity.

If φ\varphi is a linear flip then U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} acts flag transitively on Γ(n,q)\Gamma(n,q).

Proof.

Since Γ(n,q)\Gamma(n,q) is a geometry, it suffices to show that U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} acts chamber transitively. Let C=(Ci)i=1nC=(C_{i})_{i=1}^{n} and D=(Di)i=1nD=(D_{i})_{i=1}^{n} be two chambers of Γ(n,q)\Gamma(n,q). By Lemma 4.2 we can find bases {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} and {gi,hi}i=1n\{g_{i},h_{i}\}_{i=1}^{n} for CC and DD respectively such that if φ\varphi is induced by an isometry as in Lemma 3.4, both bases satisfy (a), and if φ\varphi is induced by any anti-isometry as in Lemma 3.4, both bases satisfy (b) for the same α𝔽\alpha\in\mathbb{F} of trace 0. Notice that {ei,φ(ei)}i=1n\{e_{i},\varphi(e_{i})\}_{i=1}^{n} and {gi,φ(gi)}i=1n\{g_{i},\varphi(g_{i})\}_{i=1}^{n} form bases for VV. Furthermore, the Gram matrix for β\beta is the same whether we take the basis {ei,φ(ei)}i=1n\{e_{i},\varphi(e_{i})\}_{i=1}^{n} or the basis {gi,φ(gi)}i=1n\{g_{i},\varphi(g_{i})\}_{i=1}^{n}. Similarly the Gram matrix for βφ\beta_{\varphi} does not depend on which of these two bases we consider.

Define T:VVT:V\to V by T(ei)=giT(e_{i})=g_{i} and T(φ(ei))=φ(gi)T(\varphi(e_{i}))=\varphi(g_{i}) and extend linearly. It is easy to see that TT preserves β\beta and commutes with φ\varphi, and hence also preserves βφ\beta_{\varphi}. Thus TT is an element of U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} with T(C)=DT(C)=D. ∎

Theorem 4.3.

Let φ\varphi be a linear flip of Δ\Delta.

  1. (i)

    If φ\varphi is induced by an isometry of (V,β)(V,\beta) as in Lemma 3.4 then U2n(q2)φUn(q2)×Un(q2)\mathrm{U}_{2n}(q^{2})^{\varphi}\cong\mathrm{U}_{n}(q^{2})\times\mathrm{U}_{n}(q^{2}).

  2. (ii)

    If φ\varphi is induced by an anti-isometry of (V,β)(V,\beta) as in Lemma 3.4 then U2n(q2)φGLn(q2)\mathrm{U}_{2n}(q^{2})^{\varphi}\cong\mathrm{GL}_{n}(q^{2}).

Proof.

Let {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} be a basis for VV as in Main Theorem 1A.

  1. (i)

    Define a new basis for VV by gi=ei+fig_{i}=e_{i}+f_{i} for i=1,,ni=1,\ldots,n and hi=eifih_{i}=e_{i}-f_{i} for i=1,,ni=1,\ldots,n. Order this basis as {g1,,gn,h1,,hn}\{g_{1},\ldots,g_{n},h_{1},\ldots,h_{n}\}. Direct calculation shows that with respect to this (ordered) basis, β\beta and βφ\beta_{\varphi} have Gram matrices

    M1=(2In002In) and M2=(2In002In)M_{1}=\left(\begin{array}[]{cc}2I_{n}&0\\ 0&-2I_{n}\\ \end{array}\right)\mbox{ and }M_{2}=\left(\begin{array}[]{cc}2I_{n}&0\\ 0&2I_{n}\\ \end{array}\right)

    respectively. Given a linear transformation TT of VV, we can express TT as a block matrix

    T=(ABCD).T=\left(\begin{array}[]{cc}A&B\\ C&D\\ \end{array}\right).

    Since both β\beta and βφ\beta_{\varphi} are hermitian, it follows that TT preserves both forms if and only if for i=1,2i=1,2, σ(Tt)MiT=Mi\sigma(T^{t})M_{i}T=M_{i} where TtT^{t} denotes the transpose of TT. These two requirements are by direct calculation equivalent to the following four equalities:

    σ(At)Aσ(Ct)C=σ(Dt)Dσ(Bt)B\displaystyle\sigma(A^{t})A-\sigma(C^{t})C=\sigma(D^{t})D-\sigma(B^{t})B =\displaystyle= In;\displaystyle I_{n}; (3)
    σ(At)A+σ(Ct)C=σ(Bt)B+σ(Dt)D\displaystyle\sigma(A^{t})A+\sigma(C^{t})C=\sigma(B^{t})B+\sigma(D^{t})D =\displaystyle= In;\displaystyle I_{n}; (4)
    σ(At)Bσ(Ct)D=σ(Bt)Aσ(Dt)C\displaystyle\sigma(A^{t})B-\sigma(C^{t})D=\sigma(B^{t})A-\sigma(D^{t})C =\displaystyle= 0;\displaystyle 0; (5)
    σ(At)B+σ(Ct)D=σ(Bt)A+σ(Dt)C\displaystyle\sigma(A^{t})B+\sigma(C^{t})D=\sigma(B^{t})A+\sigma(D^{t})C =\displaystyle= 0.\displaystyle 0. (6)

    Adding (3) to (4) shows that σ(At)A=σ(Dt)D=In\sigma(A^{t})A=\sigma(D^{t})D=I_{n}, and so AA and DD are unitary matrices. Adding (5) to (6) and using the fact that AA is invertible shows that B=0B=0. Similarly subtracting (5) from (6) and using the fact that DD is invertible shows that C=0C=0. Thus in fact

    T=(A00D)T=\left(\begin{array}[]{cc}A&0\\ 0&D\\ \end{array}\right)

    where AA and DD are unitary matrices. Conversely it is easy to check that if AA and DD are unitary matrices then

    (A00D)\left(\begin{array}[]{cc}A&0\\ 0&D\\ \end{array}\right)

    preserves both β\beta and βφ\beta_{\varphi}, and so lies in U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi}.

  2. (ii)

    The technique here is the same as in (i), but the details are different. We only outline this part of the proof. Define a new basis for VV by setting gi=ei+αfig_{i}=e_{i}+\alpha f_{i} for i=1,,ni=1,\ldots,n and hi=eiαfih_{i}=e_{i}-\alpha f_{i} for i=n,,ni=n,\ldots,n and order this basis {g1,,gn,h1,,hn}\{g_{1},\ldots,g_{n},h_{1},\ldots,h_{n}\}. Considering a linear transformation of VV as a block matrix TT as above, direct calculation shows that TT preserves both β\beta and βφ\beta_{\varphi} if and only if B=C=0B=C=0 and σ(At)D=In\sigma(A^{t})D=I_{n}. Conversely for any AGLn(q2)A\in\mathrm{GL}_{n}(q^{2}) the matrix

    (A00(σ(At))1)\left(\begin{array}[]{cc}A&0\\ 0&(\sigma(A^{t}))^{-1}\end{array}\right)

    preserves both β\beta and βφ\beta_{\varphi} and so lies in U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi}.∎

5 σ\sigma-Semilinear Flips

We now turn our attention to the study of σ\sigma-semilinear flips of the unitary building. In particular we prove in Theorem Main Theorem 1B: Classification of σ\sigma-Semilinear Flips that there are only 2 similarity classes of σ\sigma-semilinear flips of the unitary building.

Throughout this section, φ\varphi denotes a σ\sigma-semilinear flip. Recall that we have identified the flip with a σ\sigma-semilinear isometry fΓU(V)f\in\Gamma\mathrm{U}(V) that induces φ\varphi and satisfies f2=idf^{2}=\mathrm{id} on VV.

5.1 Geometries Induced by a σ\sigma-semilinear Flip

We’ve already seen that the geometry induced by a flip φ\varphi is related to a form βφ\beta_{\varphi} defined on VV. In the case of a linear flip we saw this form is σ\sigma-hermitian or σ\sigma-antihermitian. As we saw in Lemma 3.6, when φ\varphi is a σ\sigma-semilinear flip of the unitary building, the induced form βφ\beta_{\varphi} is symmetric.

Definition 5.1.

Let UU be a subspace of VV and let 𝒰\mathcal{U} be a basis for VV. Recall that βφ(𝒰)\beta_{\varphi}(\mathcal{U}) denotes the Gram matrix of the form βφ\beta_{\varphi} restricted to UU with respect to the basis 𝒰\mathcal{U}. The discriminant of UU is:

disc(U)={1, if det(βφ(𝒰)) is a square in 𝔽;1, if det(βφ(𝒰)) is a non-square in 𝔽;0, if U is βφ degenerate.\mathrm{disc}(U)=\left\{\begin{array}[]{r l}1,&\mbox{ if }\det(\beta_{\varphi}(\mathcal{U}))\mbox{ is a square in }\mathbb{F};\\ -1,&\mbox{ if }\det(\beta_{\varphi}(\mathcal{U}))\mbox{ is a non-square in }\mathbb{F};\\ 0,&\mbox{ if $U$ is $\beta_{\varphi}$ degenerate.}\end{array}\right.
Definition 5.2.

A square type (resp. non-square type) ii-space is an ii-dimensional subspace UU of VV with disc(U)=1\mathrm{disc}(U)=1 (resp. disc(U)=1\mathrm{disc}(U)=-1).

Lemma 5.1.

Let UU, UΓ(n,q)U^{\prime}\in\Gamma(n,q) with UUU\subseteq U^{\prime}. Let W=U,φ(U)UW=\langle U,\varphi(U)\rangle^{\perp}\cap U^{\prime}. Then

disc(U)=disc(U)disc(W).\mathrm{disc}(U^{\prime})=\mathrm{disc}(U)\mathrm{disc}(W).
Proof.

Since WW is a βφ\beta_{\varphi} orthogonal complement to UU in UU^{\prime}, we can choose a basis relative to this decomposition and so represent βφ(U)\beta_{\varphi}(U^{\prime}) as a matrix with the form

βφ(U)=(βφ(W)00βφ(U))\beta_{\varphi}(U^{\prime})=\left(\begin{array}[]{cc}\beta_{\varphi}(W)&0\\ 0&\beta_{\varphi}(U)\\ \end{array}\right)

which has determinant (detβφ(W))(detβφ(U))(\det\beta_{\varphi}(W))(\det\beta_{\varphi}(U)). ∎

We now define two pregeometries contained in Γ(n,q)\Gamma(n,q). We will prove shortly that these are in fact geometries and in Section 5.3 we show that both these geometries are flag transitive.

Definition 5.3.

Define the following pregeometries in Γ(n,q)\Gamma(n,q):

Γ1(n,q)\displaystyle\Gamma_{1}(n,q) =\displaystyle= {UΓ(n,q)|disc(U)=1 or dim(U)=n};\displaystyle\{U\in\Gamma(n,q)|\mathrm{disc}(U)=1\mbox{ or }\dim(U)=n\};
Γ1(n,q)\displaystyle\Gamma_{-1}(n,q) =\displaystyle= {UΓ(n,q)|disc(U)=1 or dim(U)=n}.\displaystyle\{U\in\Gamma(n,q)|\mathrm{disc}(U)=-1\mbox{ or }\dim(U)=n\}.
Note 5.1.

Just as in the study of Γ(n,q)\Gamma(n,q), a point of Γ1(n,q)\Gamma_{1}(n,q) (resp. Γ1(n,q)\Gamma_{-1}(n,q)) is a 1-dimensional object of Γ1(n,q)\Gamma_{1}(n,q) (resp. Γ1(n,q)\Gamma_{-1}(n,q)) and a line of Γ1(n,q)\Gamma_{1}(n,q) (resp. Γ1(n,q)\Gamma_{-1}(n,q)) is a 2-dimensional object of Γ1(n,q)\Gamma_{1}(n,q) (resp. Γ1(n,q)\Gamma_{-1}(n,q)).

It is not immediately clear why we take all nn-dimensional elements of Γ(n,q)\Gamma(n,q) in both Γ1(n,q)\Gamma_{1}(n,q) and Γ1(n,q)\Gamma_{-1}(n,q). We will see shortly (Theorem 5.11) that the nn-dimensional objects of Γ(n,q)\Gamma(n,q) all have the same βφ\beta_{\varphi} type. In order that Γ1(n,q)\Gamma_{1}(n,q) and Γ1(n,q)\Gamma_{-1}(n,q) are both geometries of rank nn, we must include these nn-dimensional objects.

We now explore the properties of the geometries Γ(n,q)\Gamma(n,q), Γ1(n,q)\Gamma_{1}(n,q) and Γ1(n,q)\Gamma_{-1}(n,q) in some more detail. Looking at these geometries will give insight into the structure of the flip. The next two results admit a uniform proof.

Lemma 5.2.

If LL is a line of Γ(n,q)\Gamma(n,q) then there are biorthogonal points uu, vv of Γ(n,q)\Gamma(n,q) so that L=u,vL=\langle u,v\rangle. If LL is a line of Γ1(n,q)\Gamma_{1}(n,q) we can assume both uu and vv are points of Γ1(n,q)\Gamma_{1}(n,q). If LL is a line of Γ1(n,q)\Gamma_{-1}(n,q) then one of uu or vv is of square type and the other is of non-square type.

Corollary 5.3.

Let LL be a line of Γ(n,q)\Gamma(n,q). Then LL contains points of both Γ1(n,q)\Gamma_{1}(n,q) and Γ1(n,q)\Gamma_{-1}(n,q).

Proof of Lemma 5.2 and Corollary 5.3.

Since LL is already β\beta totally isotropic it suffices to only consider the form βφ\beta_{\varphi}. The results follow by straightforward calculations. ∎

Lemma 5.4.

If dimU2\dim U\geq 2 and UΓ(n,q)U\in\Gamma(n,q), then UU contains points of both Γ1(n,q)\Gamma_{1}(n,q) and Γ1(n,q)\Gamma_{-1}(n,q).

Corollary 5.5.

If UΓϵ(n,q)U\in\Gamma_{\epsilon}(n,q) then UU contains a point of Γϵ(n,q)\Gamma_{\epsilon}(n,q).

Proof.

If dim(U)=1\dim(U)=1 then UU is a point of Γϵ(n,q)\Gamma_{\epsilon}(n,q). If dim(U)>1\dim(U)>1 then we can apply Lemma 5.4 to conclude that UU contains a point of Γϵ(n,q)\Gamma_{\epsilon}(n,q). ∎

Corollary 5.6.

A hh-object UU of Γ1(n,q)\Gamma_{1}(n,q) with h<nh<n has a basis {u1,,uh}\{u_{1},\ldots,u_{h}\} of square type points that are pairwise biorthogonal. Consequently any hh-object UU of Γ1(n,q)\Gamma_{1}(n,q) contains a {1,,h}\{1,\ldots,h\} flag of Γ1(n,q)\Gamma_{1}(n,q).

Proof.

The conclusion follows by induction on hh as in the proof of Corollary 3.19 with Corollary 5.5 replacing Lemma 3.14 and Lemma 5.1 replacing Lemma 3.13. ∎

Corollary 5.7.

Let MM be an nn-object of Γ(n,q)\Gamma(n,q). If MM is of square type then MM has a basis {u1,,un}\{u_{1},\ldots,u_{n}\} of pairwise biorthogonal square type points. If MM is of non-square type them MM has a basis {v1,,vn}\{v_{1},\ldots,v_{n}\} of pairwise biorthogonal points, where for i=1,,n1i=1,\ldots,n-1, viv_{i} is of square type and vnv_{n} is of non-square type.

Proof.

If MM is of square type it contains a square type point u1u_{1}. By applying Corollary 5.6 to M=Mu1M^{\prime}=M\cap\langle u_{1}\rangle^{\perp\hskip-4.26773pt\perp} we produce the remaining points. If MM is of non-square type it contains a non-square type point vnv_{n} and applying Corollary 5.6 to M=vnM^{\prime}=\langle v_{n}\rangle^{\perp\hskip-4.26773pt\perp} produces the remaining points. ∎

Remark. We will prove in Theorem 5.11 that for a fixed flip φ\varphi, every maximal object has the same βφ\beta_{\varphi} type. \lozenge

Lemma 5.8.

Let uu be a point of Γϵ(n,q)\Gamma_{\epsilon}(n,q) for ϵ{1,1}\epsilon\in\{1,-1\}. Then resΓϵ(n,q)(u)Γ1(n1,q)\mathrm{res}_{\Gamma_{\epsilon}(n,q)}(u)\cong\Gamma_{1}(n-1,q).

Proof.

This follows immediately from the proof of Corollary 3.27 once we note that if uΓ1(n,q)u\in\Gamma_{1}(n,q) then by Lemma 5.1, rur_{u} sends square type subspaces to square type subspaces, and non-square type subspaces to non-square type subspaces. Similarly if uΓ1(n,q)u\in\Gamma_{-1}(n,q) then rur_{u} sends square type subspaces to non-square type subspaces, and non-square type subspaces to square type subspaces. ∎

Theorem 5.9.

Γ1(n,q)\Gamma_{1}(n,q) and Γ1(n,q)\Gamma_{-1}(n,q) are a geometries with type and incidence inherited from Γ(n,q)\Gamma(n,q).

Proof.

The same arguments as in the proof of Theorem 3.28 work in this case, with Lemma 5.8 replacing Corollary 3.27 and Corollary 5.5 replacing Lemma 3.14. ∎

Lemma 5.10.

Let uu be a point of Γ(n,q)\Gamma(n,q) and let U=u,φ(u)U=\langle u,\varphi(u)\rangle. Then either every point of Γ(n,q)\Gamma(n,q) contained in UU is of square type, or every point is of non-square type.

Proof.

We first show that if λ\lambda has trace 0 in 𝔽\mathbb{F} then 1+λ21+\lambda^{2} lies in 𝔽q\mathbb{F}_{q} and so in particular is a square in 𝔽\mathbb{F}. Since Trσ(λ)=0\mathrm{Tr}_{\sigma}(\lambda)=0 it follows that σ(λ)=λ\sigma(\lambda)=-\lambda and so σ(λ2)=σ(λ)2=λ2\sigma(\lambda^{2})=\sigma(\lambda)^{2}=\lambda^{2}. Hence λ2𝔽q\lambda^{2}\in\mathbb{F}_{q} and thus also 1+λ2𝔽1+\lambda^{2}\in\mathbb{F}.

If uu is a point in UU, then all the other points of Γ(n,q)\Gamma(n,q) that lie in UU (except for φ(u)\varphi(u)) are of the form u+λφ(u)u+\lambda\varphi(u) for some non-zero λ\lambda of trace 0, and since these points have QφQ_{\varphi} value (1+λ2)Qφ(u)(1+\lambda^{2})Q_{\varphi}(u), we conclude that all these points have the same type as uu. Since also φ(u)\varphi(u) has the same βφ\beta_{\varphi} type as uu we conclude that all the points of UU have the same βφ\beta_{\varphi} type. ∎

Theorem 5.11.

Let MM and MM^{\prime} be nn-objects of Γ(n,q)\Gamma(n,q). Then MM and MM^{\prime} have the same βφ\beta_{\varphi} type.

Proof.

If n=1n=1 then this follows because V=u,φ(u)V=\langle u,\varphi(u)\rangle for some point uu of Γ(1,q)\Gamma(1,q). Since every point on u,φ(u)\langle u,\varphi(u)\rangle has the same βφ\beta_{\varphi} type as uu (by Lemma 5.10), when n=1n=1 the maximal objects of Γ(1,q)\Gamma(1,q) all have the same βφ\beta_{\varphi} type.

Assume n>1n>1, suppose MM is of square type and suppose MM^{\prime} is of non-square type. Then by Corollary 5.7 MM has a basis {e1,,en}\{e_{1},\ldots,e_{n}\} of biorthogonal square type points. We may then scale each eie_{i} so that βφ(ei,ei)=1\beta_{\varphi}(e_{i},e_{i})=1 for all ii. Setting fi=φ(ei)f_{i}=\varphi(e_{i}) for i=1,,ni=1,\ldots,n we obtain a β\beta hyperbolic basis 1={ei,φ(ei)}i=1n\mathcal{B}_{1}=\{e_{i},\varphi(e_{i})\}_{i=1}^{n} for VV.

Similarly by Corollary 5.7 MM^{\prime} has a basis {g1,,gn}\{g_{1},\ldots,g_{n}\} of biorthogonal points where g1,,gn1g_{1},\ldots,g_{n-1} are of square type and gng_{n} is of non-square type. After scaling we can assume that βφ(gi,gi)=1\beta_{\varphi}(g_{i},g_{i})=1 for i=1,,n1i=1,\ldots,n-1, and βφ(gn,gn)=α\beta_{\varphi}(g_{n},g_{n})=\alpha where α\alpha is a non-square in 𝔽\mathbb{F}. Let hn=α1gnh_{n}=\alpha^{-1}g_{n}. Notice that φ(hn)=σ(α1)φ(gn)\varphi(h_{n})=\sigma(\alpha^{-1})\varphi(g_{n}) and so φ(gn)=σ(α)φ(hn)\varphi(g_{n})=\sigma(\alpha)\varphi(h_{n}). It is easy to see that (hn,φ(gn))(h_{n},\varphi(g_{n})) forms a β\beta hyperbolic pair, and that βφ(hn,hn)=α1\beta_{\varphi}(h_{n},h_{n})=\alpha^{-1}, and βφ(φ(gn),φ(gn))=σ(α)\beta_{\varphi}(\varphi(g_{n}),\varphi(g_{n}))=\sigma(\alpha). This gives another β\beta hyperbolic basis, 2={g1,φ(g1),,gn1,φ(gn1),hn,σ(α)φ(hn)}\mathcal{B}_{2}=\{g_{1},\varphi(g_{1}),\ldots,g_{n-1},\varphi(g_{n-1}),h_{n},\sigma(\alpha)\varphi(h_{n})\}.

Let TT be the transition matrix from 1\mathcal{B}_{1} to 2\mathcal{B}_{2}. For i=1,2i=1,2 let BiB_{i} denote the Gram matrix of β\beta with respect to i\mathcal{B}_{i}, and let CiC_{i} denote the Gram matrix of βφ\beta_{\varphi} with respect to i\mathcal{B}_{i}. Since TT is the transition matrix from 1\mathcal{B}_{1} to 2\mathcal{B}_{2} it follows that

TB1σ(Tt)\displaystyle TB_{1}\sigma(T^{t}) =\displaystyle= B2\displaystyle B_{2} (7)
TC1Tt\displaystyle TC_{1}T^{t} =\displaystyle= C2.\displaystyle C_{2}. (8)

Since C1=I2nC_{1}=I_{2n} it follows from (8) that TTt=C2TT^{t}=C_{2}. It is easy to check that det(C2)=αq1\det(C_{2})=\alpha^{q-1} and so also det(T)2=αq1\det(T)^{2}=\alpha^{q-1}. Hence det(T)=±α(q1)/2\det(T)=\pm\alpha^{(q-1)/2} and it is easy to check that this implies Nσ(det(T))=1N_{\sigma}(\det(T))=-1.

On the other hand, since both 1\mathcal{B}_{1} and 2\mathcal{B}_{2} are β\beta hyperbolic bases, we see that B1=B2B_{1}=B_{2}, which combined with (7) forces Nσ(det(T))=1N_{\sigma}(\det(T))=1, a contradiction. Hence MM and MM^{\prime} have the same βφ\beta_{\varphi} type. ∎

Definition 5.4.

A σ\sigma-semilinear flip φ\varphi is of square type if the maximal objects of Γ(n,q)\Gamma(n,q) are of square βφ\beta_{\varphi} type. A σ\sigma-semilinear flip φ\varphi is of non-square type if the maximal objects of Γ(n,q)\Gamma(n,q) are of non-square βφ\beta_{\varphi} type.

5.2 Classification of σ\sigma-Semilinear Flips of the Unitary Building

We are now in a position to fully classify the flips of the unitary building that are induced by σ\sigma-semilinear transformations of the underlying vector space.

Main Theorem 1B: Classification of σ\sigma-Semilinear Flips.

Let φ\varphi be a σ\sigma-semilinear flip on the unitary building. Then there is a basis for VV, {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} of β\beta hyperbolic pairs so that for i=1,,n1i=1,\ldots,n-1 we have φ(ei)=fi\varphi(e_{i})=f_{i}, φ(fi)=ei\varphi(f_{i})=e_{i} and either

  1. (i)

    φ(en)=fn\varphi(e_{n})=f_{n}, φ(fn)=en\varphi(f_{n})=e_{n} or

  2. (ii)

    φ(en)=λfn\varphi(e_{n})=\lambda f_{n}, φ(fn)=σ(λ1)en\varphi(f_{n})=\sigma(\lambda^{-1})e_{n} where λ\lambda is a non-square in 𝔽\mathbb{F}.

Case (i) occurs if φ\varphi is of square type, and Case (ii) occurs if φ\varphi is of non-square type. Conversely any σ\sigma-semilinear transformation of VV that satisfies the conclusion of this theorem induces a flip of Δ\Delta.

Proof.

The forward implication follows immediately from the proof of Theorem 5.11. That is, the proof of Theorem 5.11 shows that if φ\varphi is of square type then there is a basis as described by (i), and if φ\varphi is of non-square type then there is a basis as described in (ii).

The converse follows from Lemma 3.9. ∎

This completes the proof of Main Theorem 1. It is worth noting that the geometry of β\beta isotropic βφ\beta_{\varphi} square-type subspaces induced by square-type and non-square type flips are not isomorphic, as can be seen from the fact that square-type and non-square type spaces contain different numbers of square-type points.

5.3 A Flag Transitive Automorphism Group of Γ1(n,q)\Gamma_{1}(n,q)

We now study the group of linear transformations of VV that preserve both β\beta and βφ\beta_{\varphi}. This group acts as an automorphism group of the geometry Γ(n,q)\Gamma(n,q) although it does not act flag transitively on that geometry. We prove in this section that this group acts flag transitively on both Γ1(n,q)\Gamma_{1}(n,q) and Γ1(n,q)\Gamma_{-1}(n,q).

Definition 5.5.

Let U2n(q2)φ={fU2n(q2)|βφ(u,v)=βφ(f(u),f(v)) for all u,vV}\mathrm{U}_{2n}(q^{2})^{\varphi}=\{f\in\mathrm{U}_{2n}(q^{2})|\beta_{\varphi}(u,v)=\beta_{\varphi}(f(u),f(v))\mbox{ for all }u,v\in V\}. Notice that this is the same notation we used in the case of a linear flip.

Lemma 5.12.

U2n(q2)φ=CΓU(V)(φ)U2n(q2)\mathrm{U}_{2n}(q^{2})^{\varphi}=C_{\Gamma\mathrm{U}(V)}(\varphi)\cap\mathrm{U}_{2n}(q^{2}).

Proof.

The same proof as in Lemma 4.1 holds in this case. ∎

Lemma 5.13.

Let C=(Ci)i=1nC=(C_{i})_{i=1}^{n} be a chamber of Γ1(n,q)\Gamma_{1}(n,q).

  1. (a)

    If φ\varphi is of square type then there is a basis {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} for VV with the following properties:

    1. (i)

      {ei,fi}\{e_{i},f_{i}\} is hyperbolic with respect to β\beta;

    2. (ii)

      for all i=1,,ni=1,\ldots,n, φ(ei)=fi\varphi(e_{i})=f_{i} and φ(fi)=ei\varphi(f_{i})=e_{i}; and

    3. (iii)

      for all i=1,,ni=1,\ldots,n, Ci=e1,,eiC_{i}=\langle e_{1},\ldots,e_{i}\rangle.

  2. (b)

    If φ\varphi is of non-square type then there is a basis {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} for VV with properties (i) and (iii) and

    1. (ii’)

      For i=1,,n1i=1,\ldots,n-1 φ(ei)=fi\varphi(e_{i})=f_{i} and φ(fi)=ei\varphi(f_{i})=e_{i}, but φ(en)=λfn\varphi(e_{n})=\lambda f_{n} and φ(fn)=σ(λ1)en\varphi(f_{n})=\sigma(\lambda^{-1})e_{n} where λ\lambda is any non-square in 𝔽\mathbb{F}.

Proof.

The same arguments as in the proof of Lemma 4.2 works with Corollary 5.5 replacing Lemma 3.14 and the scaling arguments from the proof of Theorem 5.11 replacing the scaling arguments from the proof of Main Theorem 1A. ∎

Lemma 5.13 is strictly stronger than Main Theorem 1B. In Main Theorem 1B we only prove that, given a maximal object we can find a basis satisfying (i) and (ii) (resp. (ii’)), what is interesting about Lemma 5.13 is that given any chamber of Δφ\Delta^{\varphi} we can additionally require that (iii) be satisfied.

Main Theorem 3: σ\sigma-Semilinear Flag Transitivity.

U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} acts flag transitively on Γ1(n,q)\Gamma_{1}(n,q).

Proof.

The same proof as in Main Theorem 2 applies here, with Lemma 5.13 replacing Lemma 4.2, but note that we must choose the λ\lambda’s to be the same if φ\varphi is a non-square type flip. ∎

The proof that U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} acts flag transitively on Γ1(n,q)\Gamma_{-1}(n,q) is similar.

We now turn to the problem of determining the isomorphism type of U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} when φ\varphi is a σ\sigma-semilinear flip of Δ\Delta. The following lemma is well known.

Lemma 5.14.

Let (W,ρ)(W,\rho) be a non-degenerate orthogonal space over a field kk. Let UU be a 2-dimensional non-degenerate subspace of WW.

  1. (i)

    If 1-1 is a square in kk, then UU is of ++ type if and only if UU is of square type.

  2. (ii)

    If 1-1 is not a square in kk, then UU is of ++ type if and only if UU is of non-square type.

Theorem 5.15.

There is a basis for VV relative to which the Gram-matrices of β\beta and βφ\beta_{\varphi} coincide.

Proof.

Choose a basis for VV as provided by Main Theorem 1. Then we have a basis {ei,fi}i=1n\{e_{i},f_{i}\}_{i=1}^{n} so that each (ei,fi)(e_{i},f_{i}) is a β\beta hyperbolic pair, and for i=1,,n1i=1,\ldots,n-1 we have

φ(ei)=fi and φ(en)=λfnφ(fi)=eiφ(fn)=σ(λ1)en\begin{array}[]{lcl}\varphi(e_{i})=f_{i}&\mbox{ and }&\varphi(e_{n})=\lambda f_{n}\\ \varphi(f_{i})=e_{i}&&\varphi(f_{n})=\sigma(\lambda^{-1})e_{n}\end{array}

If φ\varphi is a square type flip then we may assume that λ=1\lambda=1. If φ\varphi is a non-square type flip then λ\lambda is a non-square in 𝔽\mathbb{F}.

Choose a𝔽qa\in\mathbb{F}_{q} that is not a square (in 𝔽q\mathbb{F}_{q}) and choose α𝔽\alpha\in\mathbb{F} so that α2=a\alpha^{2}=a. Notice that σ(α)=α\sigma(\alpha)=-\alpha. Define a new basis as follows. For i=1,,n1i=1,\ldots,n-1 set

gi\displaystyle g_{i} =\displaystyle= ei+fi\displaystyle e_{i}+f_{i}
gi+n\displaystyle g_{i+n} =\displaystyle= α(eifi)\displaystyle\alpha(e_{i}-f_{i})
gn\displaystyle g_{n} =\displaystyle= en+λfn\displaystyle e_{n}+\lambda f_{n}
g2n\displaystyle g_{2n} =\displaystyle= α(enλfn)\displaystyle\alpha(e_{n}-\lambda f_{n})

It is easy to check that each vector gig_{i}, i=1,,2ni=1,\ldots,2n, is fixed by φ\varphi, and so since φ\varphi acts trivially on the basis {gi|i=1,,2n}\{g_{i}|i=1,\ldots,2n\} we conclude the Gram matrices of β\beta and βφ\beta_{\varphi} agree with respect to this basis. ∎

Construction 2.

For convenience, we reorder the basis {gi|i=1,,2n}\{g_{i}|i=1,\ldots,2n\} from the proof of Theorem 5.15 as follows: for i=1,,ni=1,\ldots,n set

h2i1\displaystyle h_{2i-1} =\displaystyle= gi\displaystyle g_{i}
h2i\displaystyle h_{2i} =\displaystyle= gi+n.\displaystyle g_{i+n}.

With respect to the basis {hi}i=12n\{h_{i}\}_{i=1}^{2n}, the common Gram matrix of β\beta and βφ\beta_{\varphi} is a block diagonal matrix with 2×22\times 2 blocks {Mi}i=1n\{M_{i}\}_{i=1}^{n}. If φ\varphi is of square type then for i=1,,ni=1,\ldots,n we have

Mi=(2002a).M_{i}=\left(\begin{array}[]{cc}2&0\\ 0&2a\\ \end{array}\right).

If φ\varphi is of non-square type then for i=1,,n1i=1,\ldots,n-1 we have the same matrix MiM_{i} as above and

Mn=(Trσ(λ)α(σ(λ)λ)α(σ(λ)λ)aTrσ(λ)).M_{n}=\left(\begin{array}[]{cc}\mathrm{Tr}_{\sigma}(\lambda)&\alpha(\sigma(\lambda)-\lambda)\\ \alpha(\sigma(\lambda)-\lambda)&a\mathrm{Tr}_{\sigma}(\lambda)\\ \end{array}\right).
Lemma 5.16.

Suppose γ\gamma is a non-square in 𝔽\mathbb{F}. Then Nσ(γ)N_{\sigma}(\gamma) is a non-square in 𝔽q\mathbb{F}_{q}.

Proof.

Since the norm map NσN_{\sigma} is multiplicative it suffices to show that if α\alpha is a generator of 𝔽\mathbb{F}^{\ast} then Nσ(α)N_{\sigma}(\alpha) is a non-square in 𝔽q\mathbb{F}_{q}. This follows from a straightforward calculation which shows that the square roots of Nσ(α)N_{\sigma}(\alpha) in 𝔽\mathbb{F} are not fixed by σ\sigma, and so do not lie in 𝔽q\mathbb{F}_{q}. ∎

Theorem 5.17.

Let φ\varphi be a σ\sigma-semilinear flip of Δ\Delta.

  1. (1)

    Suppose φ\varphi is a square type flip.

    1. (i)

      If nn is even or 1-1 is not a square in 𝔽q\mathbb{F}_{q} then U2n(q2)φO2n+(q)\mathrm{U}_{2n}(q^{2})^{\varphi}\cong\mathrm{O}_{2n}^{+}(q).

    2. (ii)

      If nn is odd and 1-1 is a square in 𝔽q\mathbb{F}_{q} then U2n(q2)φO2n(q)\mathrm{U}_{2n}(q^{2})^{\varphi}\cong\mathrm{O}_{2n}^{-}(q).

  2. (2)

    Suppose φ\varphi is a non-square type flip.

    1. (i)

      If nn is even or 1-1 is not a square in 𝔽q\mathbb{F}_{q} then U2n(q2)φO2n(q)\mathrm{U}_{2n}(q^{2})^{\varphi}\cong\mathrm{O}_{2n}^{-}(q).

    2. (ii)

      If nn is odd and 1-1 is a square in 𝔽q\mathbb{F}_{q} then U2n(q2)φO2n+(q)\mathrm{U}_{2n}(q^{2})^{\varphi}\cong\mathrm{O}_{2n}^{+}(q).

Proof.

Let MM denote the common Gram matrix of β\beta and βφ\beta_{\varphi} with respect to the basis {hi|i=1,,2n}\{h_{i}|i=1,\ldots,2n\} produced in Construction 2. Then, with respect to this basis, a transformation AGL(V)A\in\mathrm{GL}(V) lies in U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} if and only if both

AtMσ(A)=M and AtMA=M.A^{t}M\sigma(A)=M\mbox{ and }A^{t}MA=M.

Since MM and AA are both invertible, it follows that A=σ(A)A=\sigma(A) and so AA is defined over 𝔽q\mathbb{F}_{q}. Hence U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} consists of all qq-rational matrices in GL(V)\mathrm{GL}(V) that satisfy AtMA=M{}^{t}AMA=M.

Since the matrix MM is also defined over 𝔽q\mathbb{F}_{q}, we see that the group of matrices satisfying AtMA=MA^{t}MA=M is the full orthogonal group over 𝔽q\mathbb{F}_{q} with respect to the symmetric bilinear form whose Gram matrix is MM. So now we must determine this group. That is, U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} is isomorphic to either O2n+(q)\mathrm{O}_{2n}^{+}(q) or O2n(q)\mathrm{O}_{2n}^{-}(q).

Recall that if L1L_{1} and L2L_{2} are βφ\beta_{\varphi}-orthogonal elliptic lines, then L1L2L_{1}\perp L_{2} can be written as H1H2H_{1}\perp H_{2} where each HiH_{i} is a hyperbolic line. It follows that the isomorphism type of U2n(q2)φ\mathrm{U}_{2n}(q^{2})^{\varphi} is determined by the last two blocks Mn1M_{n-1} and MnM_{n} if nn is even, and the last block MnM_{n} if nn is odd.

The rest of the proof follows easily by combining Lemma 5.14 with the matrices from Construction 2. The only subtlety is that Lemma 5.16 is needed to show that if φ\varphi is non-square type flip then det(Mn)\det(M_{n}) is a square in 𝔽q\mathbb{F}_{q}. ∎

References

  • [1] P. Abramenko and K. Brown. Buildings: Theory and Applications, volume 248 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2008.
  • [2] C. Bennett, R. Gramlich, C. Hoffman, and S. Shpectorov. Curtis-Phan-Tits theory. In Groups, combinatorics & geometry (Durham, 2001), pages 13–29. World Sci. Publ., River Edge, NJ, 2003.
  • [3] C. Bennett, R. Gramlich, C. Hoffman, and S. Shpectorov. Odd-dimensional orthogonal groups as amalgams of unitary groups. I. General simple connectedness. J. Algebra, 312(1):426–444, 2007.
  • [4] C. Bennett and S. Shpectorov. A new proof of a theorem of Phan. J. Group Theory, 7(3):287–310, 2004.
  • [5] B. Carr. On Flips of Unitary Buildings. PhD thesis, Bowling Green State University, 2010.
  • [6] B. Carr. On flips of unitary buildings II: Geometries related to flips. preprint, 2010.
  • [7] R. Gramlich, C. Hoffman, W. Nickel, and S. Shpectorov. Even-dimensional orthogonal groups as amalgams of unitary groups. J. Algebra, 284(1):141–173, 2005.
  • [8] R. Gramlich, M. Horn, and B. Mühlherr. Abstract involutions of algebraic groups and of kac-moody groups. Preprint (2008). http://arxiv.org/abs/0905.1280.
  • [9] M. Horn. Involutions of Kac-Moody Groups. PhD thesis, Technischen Universität Darmstadt, 2009.
  • [10] I. M. Isaacs. Character Theory of Finite Groups. American Mathematical Society, New York, 2006.
  • [11] A. Pasini. Diagram geometries. Oxford Science Publications. The Clarendon Press Oxford University Press, New York, 1994.
  • [12] D. Taylor. The geometry of the classical groups, volume 9 of Sigma Series in Pure Mathematics. Heldermann Verlag, Berlin, 1992.
  • [13] J. Tits. Buildings of spherical type and finite BN-pairs. Springer-Verlag, Berlin, 1974. Lecture Notes in Mathematics, Vol. 386.
  • [14] J. Tits. Ensembles ordonnés, immeubles et sommes amalgamées. Bull. Soc. Math. Belg. Sér. A, 38:367–387 (1987), 1986.