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

Fixed points for group actions on
22-dimensional affine buildings

Jeroen Schillewaert, Koen Struyve and Anne Thomas Jeroen Schillewaert, Department of Mathematics, University of Auckland, 38 Princes Street, 1010 Auckland, New Zealand
Koen Struyve, Burgemeester Couckestraat 4, 8720 Dentergem, Belgium
Anne Thomas, School of Mathematics & Statistics, Carslaw Building F07, University of Sydney NSW 2006, Australia
[email protected], [email protected], [email protected]
Abstract.

We prove a local-to-global result for fixed points of finitely generated groups acting on 2-dimensional affine buildings of types A~2\tilde{A}_{2} and C~2\tilde{C}_{2}. Our proofs combine building-theoretic arguments with standard results for CAT(0)\operatorname{CAT}(0) spaces.

This research of the first author is supported by the New Zealand Marsden fund through grant UOA-2122 and that of the third author by ARC grant DP180102437.

1. Introduction

The study of local-to-global results for fixed points of groups acting on affine buildings originated with Serre, who proved such a result for simplicial trees [26, Corollary 3 of Section 6.5] and introduced property (FA). Serre’s result was extended by Morgan and Shalen to \mathbb{R}-trees [16, Proposition II.2.15]. We prove a similar result for certain 2-dimensional affine buildings, possibly nondiscrete. We note that the results of Serre and Morgan–Shalen immediately imply the following theorem for buildings of type A~1×A~1\tilde{A}_{1}\times\tilde{A}_{1}, since such buildings are products of trees.

Theorem A.

Let GG be a finitely generated group of automorphisms of a 2-dimensional affine building XX of type A~2\tilde{A}_{2} or C~2\tilde{C}_{2} (possibly nondiscrete). If every element of GG fixes a point of XX, then GG fixes a point of XX.

Remark 1.1.

If a 22-dimensional affine building XX is the Bruhat–Tits building for a reductive group over a field with valuation, or XX is discrete, then XX has type A~2\tilde{A}_{2}, C~2\tilde{C}_{2} or G~2\tilde{G}_{2}. We give a detailed explanation of why our techniques do not extend to type G~2\tilde{G}_{2} in Remarks 2.3 and 3.16.

By applying Theorem A to all finitely generated subgroups and then invoking a result of Caprace and Lytchak [6, Theorem 1.1], we obtain the following corollary for non-finitely generated groups.

Corollary 1.2.

Suppose a group GG acts on a complete 2-dimensional affine building XX of type A~2\tilde{A}_{2} or C~2\tilde{C}_{2} such that every element of GG fixes a point of XX. Then GG fixes a point in the bordification X¯=XX\overline{X}=X\cup\partial X of XX.

When XX is discrete, Corollary 1.2 confirms Conjecture 1.2 of Marquis [13] for this class of buildings. In this paper Marquis introduces Property (FB): every measurable action of a group GG by type-preserving simplicial isometries on a finite rank discrete building stabilises a spherical residue. Since continuous actions are measurable, we see that for discrete groups this is a higher-rank analogue of Serre’s property (FA). As explained in [13, Remark 3.4], Corollary 1.2 combined with the Morgan–Shalen result for trees implies a special case of Conjecture 1.1 of [13].

Corollary 1.3.

Let GG be an almost connected locally compact group which acts measurably by type-preserving simplicial isometries on a discrete 2-dimensional affine building of type A~2\tilde{A}_{2} or C~2\tilde{C}_{2}. Then GG fixes a point of XX.

Many other local-to-global results similar to Theorem A are known. Parreau in [21, Corollaire 3] proved a similar result for subgroups Γ\Gamma of connected reductive groups 𝒢\mathcal{G} over certain fields FF, where Γ\Gamma is generated by a bounded subset of 𝒢(F)\mathcal{G}(F) and the action is on the completion of the associated Bruhat–Tits building. Breuillard and Fujiwara established a quantitative version of Parreau’s result for discrete Bruhat–Tits buildings [3, Theorem 7.16] and asked whether their result holds for the isometry group of an arbitrary affine building. Leder and Varghese in [11], using work of Sageev [25], obtained a similar result for groups acting on finite-dimensional CAT(0)\operatorname{CAT}(0) cube complexes. However, a statement similar to Theorem A is false for infinite-dimensional CAT(0)\operatorname{CAT}(0) cubical complexes, as shown by Osajda using actions of infinite free Burnside groups [19].

While we were preparing this paper, Norin, Osajda and Przytycki [18] proved a result closely related to Theorem A.

Theorem 1.4 (Theorem 1.1 of [18]).

Let XX be a CAT(0)\operatorname{CAT}(0) triangle complex and let GG be a finitely generated group acting on XX with no global fixed point. Assume that either each element of GG fixing a point of XX has finite order, or XX is locally finite, or XX has rational angles. Then GG has an element with no fixed point in XX.

Remark 1.5.

One of the most important classes of CAT(0)\operatorname{CAT}(0) triangle complexes is that of 22-dimensional discrete affine buildings, and since such buildings have rational angles, Theorem 1.4 applies to these. In particular, it applies to discrete buildings of type G~2\tilde{G}_{2}. On the other hand, Theorem A is valid for both discrete and nondiscrete buildings (of types A~2\tilde{A}_{2} and C~2\tilde{C}_{2}), and as explained below, some parts of our argument are straightforward in the discrete case. Our proof combines general CAT(0)\operatorname{CAT}(0)-space techniques with arguments specific to spherical and affine buildings, while [18] uses Helly’s theorem from [8] together with sophisticated results including Masur’s theorem on periodic trajectories in rational billiards [15], and Ballmann and Brin’s methods for finding closed geodesics in 22-dimensional locally CAT(0)\operatorname{CAT}(0) complexes [2].

The next result follows from Theorem A together with the fact that every isometry of a complete affine building is semisimple, that is, either it fixes a point or it is hyperbolic [20, Corollaire 4.2]. (The corresponding consequence of Theorem 1.4 is [18, Corollary 1.3].)

Corollary 1.6.

If a finitely generated group GG acts without a fixed point on a complete 2-dimensional (possibly nondiscrete) affine building XX of type A~2\tilde{A}_{2} or C~2\tilde{C}_{2}, then GG contains a hyperbolic isometry, in particular G\mathbb{Z}\leq G.

We note that Swenson proved in [28, Theorem 11] that if a group GG acts properly discontinuously and cocompactly on an (unbounded, proper) CAT(0)\operatorname{CAT}(0) space XX, then GG has an element of infinite order. So Corollary 1.6 can be viewed as a strengthening of this result in some special cases.

Corollary 1.6 in particular gives a negative answer to the question of whether finitely generated infinite torsion groups can act on discrete 22-dimensional affine buildings (of types A~2\tilde{A}_{2} or C~2\tilde{C}_{2}) without fixing a point. This question formed the initial motivation for our work, and was generously suggested to the first author by Pierre-Emmanuel Caprace as a test case to complete a first (small) step towards a Tits Alternative for groups acting on CAT(0)\operatorname{CAT}(0) spaces.

We prove Theorem A in Section 3 by first establishing a result (Corollary 3.4) concerning the the realisation of distances between fixed point sets of finitely generated subgroups of GG; this result is immediate for XX discrete. The proof of Corollary 3.4 given here uses a theorem of Caprace and Lytchak [6] together with properties of panel-trees. (Recently, Leung, Schillewaert and Thomas [12] gave an alternative proof of Corollary 3.4.) We then reduce to the case that XX is a metrically complete \mathbb{R}-building and the action of GG is type-preserving. We remark that any discrete building is metrically complete, but a nondiscrete building is not necessarily metrically complete, even when it is a Bruhat–Tits building [14]. Moreover, the Cauchy completion of a nondiscrete building may not be a building at all [10]. In order to apply results for complete CAT(0)\operatorname{CAT}(0) spaces as well as for buildings, we use properties of ultrapowers of affine buildings due to Kleiner and Leeb [9] and Struyve [27].

We next show that if GG has two finitely generated proper subgroups whose fixed point sets are nonempty and disjoint, then GG contains a hyperbolic element (see Proposition 3.8). Theorem A is obtained by combining this result with an easy induction on the number of generators of GG. To prove Proposition 3.8, we construct an orbit of an element gGg\in G together with a point ξX\xi\in\partial X such that the Busemann function with respect to ξ\xi is unbounded on this orbit, hence gg is hyperbolic. A key role in this construction is played by the “local lemmas” which we establish in Section 2. These guarantee that for Δ\Delta a spherical building which occurs as the link of a vertex in XX, for any point in Δ\Delta there exists an element gΔg_{\Delta} of GG which acts on Δ\Delta and maps this point “far away” from itself. The element gg is then the product of two gΔg_{\Delta} which are chosen carefully in relation to ξ\xi. The results of Section 2 for spherical buildings are proved using the description of Δ\Delta as a point-line geometry (which varies by type). Our arguments in Section 3 are partly type-free.

Throughout the paper, we assume knowledge of discrete buildings on the level of the references Abramenko–Brown [1] or Ronan [23]. Our main reference for nondiscrete affine buildings is Parreau’s work [20] (a translation of which into English is available at [22]). We use many results from [20], and mostly follow its terminology and notation. We also assume basic knowledge of CAT(0)\operatorname{CAT}(0) spaces, see e.g. Bridson–Haefliger [4].

Acknowledgements We thank Martin Bridson, Pierre-Emmanuel Caprace, Timothée Marquis, Dave Witte Morris, Damian Osajda and James Parkinson for helpful conversations, and an anonymous referee for careful reading, including realising the necessity of Corollary 3.4 in the nondiscrete case. We also thank the University of Auckland for supporting travel between the first and third authors by means of an FRDF grant and a PBRF grant.

2. Local lemmas

In this section we consider the action of a group GG on a spherical building Δ\Delta of type A2A_{2} or C2C_{2}. In Section 3, Δ\Delta will be the link of a vertex of XX, hence we refer to the results below as local lemmas.

We realise Δ\Delta as a CAT(1)\operatorname{CAT}(1) space, so that the distances referred to in each statement are the distances in this metric on Δ\Delta. In particular, opposite points of Δ\Delta are at distance π\pi. For Δ\Delta of type C2C_{2} (respectively, A2A_{2}) we consider Δ\Delta as a generalised quadrangle (respectively, a projective plane). For g,hGg,h\in G and pp a panel (point or line) of Δ\Delta, we then write pgp^{g} for the panel obtained by acting on pp by gg, and put pgh:=(pg)hp^{gh}:=(p^{g})^{h}.

Our local lemma in type C2C_{2} is as follows.

Lemma 2.1.

Let Δ\Delta be a building of type C2C_{2} (realised as a CAT(1)\operatorname{CAT}(1)-space) and let GG be a group of type-preserving automorphisms of Δ\Delta. If xx is a point of Δ\Delta (not necessarily a panel) and pp is a panel of Δ\Delta at minimum distance from xx, then at least one of the following holds:

  1. (1)

    There is an element gGg\in G mapping pp to a panel opposite pp.

  2. (2)

    There is a panel pp^{\prime} of Δ\Delta which is fixed by GG such that d(p,x)<π2d(p^{\prime},x)<\frac{\pi}{2}.

Moreover in case (1), d(p,gx)7π8d(p,gx)\geq\frac{7\pi}{8}.

Proof.

By duality, we may assume that the panel pp is a line ll. As d(p,x)π8d(p,x)\leq\frac{\pi}{8}, in case (1) we obtain that d(p,gx)7π8d(p,gx)\geq\frac{7\pi}{8}. Now assume we are not in case (1). Then lgll^{g}\cap l\neq\emptyset for all gGg\in G. If lg=ll^{g}=l for all gGg\in G then (2) holds with p=lp^{\prime}=l. So suppose lg0ll^{g_{0}}\neq l for some g0Gg_{0}\in G, and let q=llg0q=l\cap l^{g_{0}}. If qlhq\notin l^{h} for some hG{g0}h\in G\setminus\{g_{0}\}, then lhlg0=l^{h}\cap l^{g_{0}}=\emptyset (otherwise there is a triangle). But then lg0h1l=l^{g_{0}h^{-1}}\cap l=\emptyset, a contradiction. So q=gGlgq=\bigcap_{g\in G}l^{g}, and hence qg=qq^{g}=q for all gGg\in G. Let p=qp^{\prime}=q, then since qlq\in l and p=lp=l we get d(p,p)=π/4d(p,p^{\prime})=\pi/4. Since d(x,p)π/8d(x,p)\leq\pi/8, we conclude d(x,p)3π8<π/2d(x,p^{\prime})\leq\frac{3\pi}{8}<\pi/2. ∎

A different statement is required in type A2A_{2}, where opposite panels have distinct types.

Lemma 2.2.

Let Δ\Delta be a building of type A2A_{2} (realised as a CAT(1)-space) and GG be a group of type-preserving automorphisms of Δ\Delta. If xx is a point of Δ\Delta (not necessarily a panel), cc is a chamber of Δ\Delta containing xx and pp is a panel of cc, then at least one of the following holds:

  1. (1)

    There exists gGg\in G mapping cc to a chamber which contains a panel opposite pp.

  2. (2)

    There is a panel pp^{\prime} of Δ\Delta which is fixed by GG such that d(p,x)<π2d(p^{\prime},x)<\frac{\pi}{2}.

Moreover in case (1), d(p,gx)2π3d(p,gx)\geq\frac{2\pi}{3}.

Proof.

By duality and abuse of notation we may let cc be the incident point-line pair (p,l)(p,l). Assume first that there is a gGg\in G such that plgp\notin l^{g}. Then lgl^{g} is opposite pp, and (1) holds. As GG is type-preserving and lgl^{g} is opposite pp, d(p,gx)2π3d(p,gx)\geq\frac{2\pi}{3}, as required. Suppose now that pgGlgp\in\bigcap_{g\in G}l^{g}. If lg=ll^{g}=l for all gGg\in G then (2) holds, so assume there is a gGg\in G such that p=lglp=l^{g}\cap l. Then ph=lghlh=pp^{h}=l^{gh}\cap l^{h}=p for all hGh\in G, and (2) holds. ∎

Remark 2.3.

In type G2G_{2}, opposite panels again have the same type. However, we cannot expect a local lemma of the same form as Lemma 2.1. For example, let Δ\Delta be a thin building of type G2G_{2}. This has CAT(1)\operatorname{CAT}(1) realisation the subdivision of the unit circle into 1212 arcs, each of length π6\frac{\pi}{6}. Let GG be the group of type-preserving automorphisms of Δ\Delta generated by a rotation gg through angle 2π3\frac{2\pi}{3}. Then for every panel pp of Δ\Delta, we have d(p,gp)=d(p,g2p)=2π3d(p,gp)=d(p,g^{2}p)=\frac{2\pi}{3}. Since no panel is mapped to an opposite by any element of GG, and GG does not fix any panels, neither (1) nor (2) of Lemma 2.1 holds in type G2G_{2}.

The best “local lemma” that we have been able to establish in type G2G_{2} is the following trichotomy. We omit its proof, which is considerably more involved than the proofs of Lemmas 2.1 and 2.2 above, since we will not actually use this result. We continue our explanation of why our techniques do not extend to affine buildings of type G~2\tilde{G}_{2} in Remark 3.16.

Lemma 2.4.

Let Δ\Delta be a building of type G2G_{2} (realised as a CAT(1)\operatorname{CAT}(1)-space) and let GG be a group of type-preserving automorphisms of Δ\Delta. If xx is a point of Δ\Delta (not necessarily a panel) and pp is a panel of Δ\Delta at minimum distance from xx, then at least one of the following three possibilities must hold:

  1. (1)

    There is an element gGg\in G mapping pp to a panel opposite pp.

  2. (2)

    There are elements g,hGg,h\in G such that pp, gpgp and hphp lie in a common apartment of Δ\Delta, and d(p,gp)=d(gp,hp)=d(hp,p)=2π3d(p,gp)=d(gp,hp)=d(hp,p)=\frac{2\pi}{3}.

  3. (3)

    There is a panel pp^{\prime} of Δ\Delta which is fixed by GG such that d(p,x)<π2d(p^{\prime},x)<\frac{\pi}{2}.

Moreover in cases (1) and (2), d(p,gx)7π12d(p,gx)\geq\frac{7\pi}{12}.

3. Proof of the main theorem

Let XX be an affine building, defined as in [20, Section 1.2], of type A~2\tilde{A}_{2} or C~2\tilde{C}_{2}. We equip XX with the maximal system of apartments. Each apartment of XX is modelled on the pair (𝔸,W)(\mathbb{A},W), where 𝔸\mathbb{A} is a 22-dimensional real vector space and WW is a subgroup of the affine isometry group of 𝔸\mathbb{A} such that the linear part of WW is a finite reflection group W¯\overline{W} of type A2A_{2} or C2C_{2}, respectively.

Now as in [20, Section 1.3.2], each facet of XX has a type given by the type of the corresponding facet of the fundamental Weyl chamber. In particular, the codimension one facets of XX have just 22 possible types, with one facet of each of these types bounding each sector (Weyl chamber) in XX, and the types of facets in any given apartment are invariant under translations of this apartment. In type C~2\tilde{C}_{2}, each wall of XX thus also has a well-defined type, induced by the type of the (opposite) facets it contains, and parallel walls in the same apartment have the same type. We remark that this concept of type is different to the usual definition of type for discrete buildings, where a discrete building of type a rank nn Coxeter system has nn distinct types of panels.

As in [24, Section 6.8], we define an automorphism of XX to be an isometry of XX which maps facets to facets and apartments to apartments. We use this definition rather than the notion of automorphism from [20, Definition 2.5], since by [20, Proposition 2.5] the latter is necessarily type-preserving, and we do not wish to impose this restriction.

Now let GG be a finitely generated group of automorphisms of XX such that every element of GG fixes a point of XX. In Section 3.1 we prove a result, Corollary 3.4, concerning the realisation of distances between fixed point sets of finitely generated subgroups of GG. We next establish several reductions, in Section 3.2. Then in Section 3.3, assuming GG has two proper finitely generated subgroups whose fixed point sets are nonempty and disjoint, we construct an element gGg\in G, a sequence of points {ai}\{a_{i}\} in XX with a2k=gka1a_{2k}=g^{k}a_{1} for all k1k\geq 1, and a point ξ\xi in X\partial X, the visual boundary of XX. In Sections 3.4 and 3.5 we show, using these constructions and Busemann functions, that the sequence {ai}\{a_{i}\} is unbounded, hence gg must be hyperbolic, a contradiction. As explained in Section 3.3, combining this with an induction completes the proof of Theorem A.

3.1. Distance between fixed point sets

The main result in this section is Corollary 3.4, which will be essential to our constructions in Section 3.3.

The following lemma and its proof were kindly provided to us by a referee, and shorten our initial argument for Corollary 3.4.

Lemma 3.1.

Let YY be a complete finite-dimensional CAT(0)\operatorname{CAT}(0) space and let AA and BB be two nonempty closed convex subsets. Let d=d(A,B)d=d(A,B). Then either there exist points aA,bBa\in A,b\in B such that d(a,b)=d=d(A,B)d(a,b)=d=d(A,B), or there exists ξAB\xi\in\partial A\cap\partial B.

Proof.

For any R0R\geq 0, the RR-neighbourhood 𝒩R(B)={yYd(y,B)R}\mathcal{N}_{R}(B)=\{y\in Y\mid d(y,B)\leq R\} is a closed convex subset of YY. Therefore setting An=A𝒩d+1n(B)A_{n}=A\cap\mathcal{N}_{d+\frac{1}{n}}(B) for all integers n>0n>0, we obtain a descending chain of closed convex subsets of AA. If the intersection nAn\cap_{n}A_{n} is nonempty, then any point α\alpha contained in it satisfies d(α,B)=d=d(A,B)d(\alpha,B)=d=d(A,B). Otherwise, [6, Theorem 1.1] ensures the existence of a point ξnAn\xi\in\cap_{n}\partial A_{n}. Now the geodesic ray [a1,ξ)[a_{1},\xi) emanating from any point a1A1a_{1}\in A_{1} is at bounded distance from BB by the definition of A1A_{1}, and so in particular we have ξAB\xi\in\partial A\cap\partial B. ∎

We now set up some further background needed for the proof of Corollary 3.4. The following result is likely known to experts, but we could not find it stated explicitly in the literature.

Lemma 3.2.

The fixed point set of a finitely generated group of automorphisms of XX intersects each apartment of XX in a finite intersection of closed half-apartments.

Proof.

Since the group is finitely generated it suffices to prove the result for an individual automorphism gg of XX. Let 𝒜\mathcal{A} be an apartment of XX. Then by [24, Proposition 9.1] (see also [21, Proposition 2.14]), I:=𝒜g(𝒜)I:=\mathcal{A}\cap g(\mathcal{A}) is a finite intersection of closed half-apartments bounded by walls, and moreover gg acts like an element of WW on II. Now the fixed point set of any element of WW acting on an entire apartment is also a finite intersection of closed half-apartments. Therefore the intersection of this fixed point set with II is also such a finite intersection, proving the lemma. ∎

Given a panel pp_{\infty} of the spherical building at infinity of XX, the corresponding panel tree T=T(p)T=T(p_{\infty}) (see, for example, [29]) has as points the equivalence classes of sector panels belonging to the same parallelism class pp_{\infty}, where the equivalence relation \mathcal{R} is defined by intersecting in a half-line (\mathcal{R} is transitive since CAT(0)\operatorname{CAT}(0)-spaces are uniquely geodesic [4, Proposition II.1.4]). Let tat_{a} and tbt_{b} be two points of TT and let rar_{a} and rbr_{b} be representatives of tat_{a} and tbt_{b}, respectively. By [17, Lemma 4.1] there exists an apartment containing subrays rar_{a}^{\prime} of rar_{a} and rbr_{b}^{\prime} of rbr_{b}. The distance d(ta,tb)d(t_{a},t_{b}) is defined to be the distance between the parallel halflines rar_{a}^{\prime} and rbr_{b}^{\prime}. This definition is clearly independent of the choices made.

Let ψ\psi be the map which sends a sector panel to its equivalence class with respect to \mathcal{R}. We define a map ϕ:XT\phi:X\to T as follows. Given a point xXx\in X, let SxS_{x} be the (unique) sector panel through xx which has pp_{\infty} in its boundary, and define ϕ(x)=ψ(Sx)\phi(x)=\psi(S_{x}).

If a group GG acts on XX and fixes pp_{\infty} then it has a naturally induced action on TT. Indeed, for tTt\in T we define g(t)g(t) to be the equivalence class of sector panels which is the image under gg of the equivalence class of sector panels corresponding to tt.

Lemma 3.3.

Let GG be a finitely generated group of elliptic automorphisms of XX fixing pp_{\infty}. Then the fixed point set in TT of GG is the image in TT under ϕ\phi of the fixed point set in XX of GG.

Proof.

If xXx\in X is fixed by GG then since GG also fixes pp_{\infty} it fixes the (unique) sector panel through xx which has pp_{\infty} in its boundary. Hence by definition ϕ(x)\phi(x) is fixed by GG. Conversely, if tTt\in T is fixed by GG consider a half-line LL belonging to a sector panel in the pre-image ψ1(t)\psi^{-1}(t) and gGg\in G. Then LgLL^{g}\cap L is a half-line, say MgM_{g}. Since gg is elliptic MgM_{g} is pointwise fixed by gg, as otherwise either gg or g1g^{-1} would map MgM_{g} into a proper subset of itself, thus gg would act as a translation with axis containing MgM_{g} and hence be hyperbolic. Now if G=g1,,gnG=\langle g_{1},\dots,g_{n}\rangle then the half-line M=i=1nMgiM=\cap_{i=1}^{n}M_{g_{i}} is fixed pointwise by GG. Let xx be a point on MM, then ϕ(x)=t\phi(x)=t and the lemma is proved. ∎

By [21, Corollary 2.19], since we are working with the maximal system of apartments, the CAT(0)\operatorname{CAT}(0) boundary of XX is the geometric realisation of its spherical building at infinity.

Corollary 3.4.

Let GG be a finitely generated group of type-preserving automorphisms of a complete 2-dimensional Euclidean building XX. If A:=Fix(GA)A:=\operatorname{Fix}(G_{A}) and B:=Fix(GB)B:=\operatorname{Fix}(G_{B}) are two nonempty fixed point sets of finitely generated subgroups GAG_{A} and GBG_{B}, then there exist points αA\alpha^{\star}\in A and βB\beta^{\star}\in B such that d(α,β)=d(A,B)d(\alpha^{\star},\beta^{\star})=d(A,B).

Proof.

By Lemma 3.1 we may assume that there exists ξAB\xi\in\partial A\cap\partial B. If ξ\xi is not a panel, by Lemma 3.2 we may consider sectors SaAS_{a}\subseteq A and SbBS_{b}\subseteq B both containing ξ\xi in their boundary and based at aAa\in A and bBb\in B respectively. By [17, Lemma 4.1] there exists an apartment containing subsectors of SaS_{a} and SbS_{b} and hence SaSbS_{a}\cap S_{b}\neq\emptyset, thus ABA\cap B\neq\emptyset. Suppose thus that ξ\xi is a panel. By [7, 1.3] the subsets TAT_{A} and TBT_{B} of the panel tree TT corresponding to ξ\xi which are fixed by AA and BB respectively are disjoint nonempty closed subtrees of TT. Hence by [7, 1.1] there is a unique shortest geodesic in TT having its initial point tat_{a} in TAT_{A} and its terminal point tbt_{b} in TBT_{B}. Note that by Lemma 3.3 and [4, Proposition II.2.2] no points in AA and BB can be at distance less than d(ta,tb)d(t_{a},t_{b}). By definition of distance in the panel tree d(A,B)=d(ta,tb)d(A,B)=d(t_{a},t_{b}) and the proof is complete.∎

3.2. Reductions

Lemma 3.5.

We may assume XX is an \mathbb{R}-building in which each point is a special vertex.

Proof.

By the last paragraph in the Remarques on [20, p. 6] (see also [24, Remark 6.3(d)]), we may regard XX as an affine building in which each apartment is modelled on the pair (𝔸,W~)(\mathbb{A},\widetilde{W}), where W~\widetilde{W} is the group of all affine isometries of 𝔸\mathbb{A} whose linear part is W¯\overline{W}. Hence we may assume that XX is an \mathbb{R}-building in which every point is a special vertex. We note that if a point xXx\in X was not originally a special vertex, then the link of xx in this \mathbb{R}-building structure will be a spherical building in which at least some of the panels are only contained in 22 chambers. Thus in general the \mathbb{R}-building XX will not be thick. ∎

Lemma 3.6.

We may pass to the ultrapower of XX, hence we may in particular assume that XX is metrically complete.

Proof.

By [27, Lemma 4.4] (which uses results from [9]) the \mathbb{R}-building XX can be isometrically embedded in a metrically complete \mathbb{R}-building XX^{\prime}, the ultrapower of XX. Moreover XX^{\prime} has the same type as XX, and the GG-action on XX extends to XX^{\prime}. Suppose GG fixes a point of XX^{\prime}. Then all GG-orbits on XX are bounded. Hence as GG is finitely generated, by [27, Main Result 1] GG fixes a point of XX. ∎

Lemma 3.7.

Let GG be a group of automorphisms of a metrically complete affine building XX. If its type-preserving subgroup GG^{\prime} fixes a point of XX, then GG fixes a point of XX.

Proof.

Assume GG^{\prime} fixes xXx\in X. Since [G:G][G:G^{\prime}] is finite, the GG-orbit of xx is bounded, hence GG fixes a point by the Bruhat–Tits fixed point theorem [5, Proposition 3.2.4]. ∎

3.3. Constructions

By the results of Section 3.2, we may assume from now on that XX is a metrically complete \mathbb{R}-building in which each point is a special vertex, and that the action of GG on XX is type-preserving.

Proposition 3.8.

Suppose GG has two proper finitely generated subgroups G0G_{0} and G1G_{1} such that the respective fixed point sets B0:=Fix(G0)B_{0}:=\operatorname{Fix}(G_{0}) and B1:=Fix(G1)B_{1}:=\operatorname{Fix}(G_{1}) are nonempty and disjoint. Then GG contains a hyperbolic element.

Proof of Theorem A, assuming the result of Proposition 3.8.

The group GG is finitely generated, with say G=s1,,snG=\langle s_{1},\dots,s_{n}\rangle. By hypothesis, each si\langle s_{i}\rangle has a nonempty fixed set. Then an induction with G0=s1,,siG_{0}=\langle s_{1},\dots,s_{i}\rangle and G1=si+1G_{1}=\langle s_{i+1}\rangle completes the proof of Theorem A. ∎

The remainder of the paper is devoted to the proof of Proposition 3.8. We start with a general result which will be used many times.

Lemma 3.9.

Let 𝐫=[a,ξ)\mathbf{r}=[a,\xi) and 𝐫=[a,ξ)\mathbf{r^{\prime}}=[a^{\prime},\xi) be geodesic rays in XX which have the same endpoint. If 𝐫\mathbf{r} is contained in a wall then so is 𝐫\mathbf{r}^{\prime}. Moreover, if XX is of type C~2\tilde{C}_{2}, then the walls containing 𝐫\mathbf{r} and 𝐫\mathbf{r^{\prime}} are of the same type.

Proof.

Consider a chamber CC in the spherical building at infinity containing the common endpoint ξ\xi of 𝐫\mathbf{r} and 𝐫\mathbf{r^{\prime}}. Let SS and SS^{\prime} be the sectors determined by CC and respectively aa and aa^{\prime} (that is, SS is the union of all geodesic rays based at aa with endpoint lying in CC, and similarly for SS^{\prime} and aa^{\prime}). Then SS and SS^{\prime} are asymptotic, so by [21, Corollary 1.6], the sectors SS and SS^{\prime} have a common subsector S′′S^{\prime\prime}. Moreover, S′′S^{\prime\prime} contains a geodesic ray 𝐫′′\mathbf{r^{\prime\prime}} which is parallel to both 𝐫\mathbf{r} and 𝐫\mathbf{r^{\prime}}, since S′′S^{\prime\prime} is contained in both SS and SS^{\prime} and has ξ\xi in its boundary. Since being parallel in an apartment preserves the property of being contained in a wall the first statement of the lemma is proved. The second statement holds in type C~2\tilde{C}_{2} because 𝐫\mathbf{r} and 𝐫\mathbf{r^{\prime}} are contained in parallel walls. ∎

We now define an angle α\alpha to equal 7π8\frac{7\pi}{8} or 2π3\frac{2\pi}{3} as XX is of type C~2\tilde{C}_{2} or A~2\tilde{A}_{2}, respectively (that is, α\alpha is the angle appearing in the corresponding local lemma). In the remainder of this section, we will construct an element gGg\in G and a point ξ\xi in X\partial X, the visual boundary of XX, together with a sequence of points {ai}i=0\{a_{i}\}_{i=0}^{\infty} of XX such that:

Lemma 3.10.

For all i1i\geq 1, we have ai(ξ,ai+1)α\angle_{a_{i}}(\xi,a_{i+1})\geq\alpha.

We prove Lemma 3.10 by induction, with i=0,1,2i=0,1,2 handled separately first. The case i=1i=1 includes the construction of ξ\xi, and gg will be the product of elements of GG which appear for i=1,2i=1,2. A schematic for our constructions in type C~2\tilde{C}_{2} is given in Figure 1, where we sketch the geometric situation in a thin building (that is, the Euclidean plane). The delicate part of the proof is to show, using apartments and retractions, that key portions of this sketch “lift” to the affine building XX. We give further figures in both types below.

\begin{overpic}[width=390.25534pt]{schematic-C2} \put(-2.0,2.0){$\xi$} \put(-2.0,9.5){$\xi$} \put(14.0,4.0){$B_{0}$} \put(52.0,11.0){$B_{1}$} \put(21.5,3.5){\footnotesize{$a_{0}$}} \put(41.5,11.0){\footnotesize{$a_{1}$}} \put(35.0,11.5){\footnotesize{$p_{1}$}} \put(46.5,11.5){\footnotesize{$g_{1}p_{1}$}} \put(55.0,1.5){\footnotesize{$a_{2}$}} \put(47.5,1.0){\footnotesize{$p_{2}$}} \put(61.0,1.0){\footnotesize{$g_{2}p_{2}$}} \put(72.5,10.0){\footnotesize{$a_{3}$}} \put(66.0,10.5){\footnotesize{$p_{3}$}} \put(78.5,10.5){\footnotesize{$g_{3}p_{3}$}} \put(90.0,1.5){\footnotesize{$a_{4}$}} \put(83.5,1.0){\footnotesize{$p_{4}$}} \put(96.0,1.0){\footnotesize{$g_{4}p_{4}$}} \put(40.5,6.0){\footnotesize{$g_{1}$}} \put(55.0,6.0){\footnotesize{$g_{2}$}} \put(72.0,6.0){\footnotesize{$g_{3}$}} \put(90.0,6.0){\footnotesize{$g_{4}$}} \end{overpic}
Figure 1. A schematic for our constructions in type C~2\tilde{C}_{2}. Here, we will define g=g2g1g=g_{2}g_{1}, and ai=gai2a_{i}=ga_{i-2} and gi=ggi2g1g_{i}=gg_{i-2}g^{-1} for all i3i\geq 3.

Now, as in the statement of Proposition 3.8, assume that GG has two proper finitely generated subgroups G0G_{0} and G1G_{1} such that the respective fixed point sets B0:=Fix(G0)B_{0}:=\operatorname{Fix}(G_{0}) and B1:=Fix(G1)B_{1}:=\operatorname{Fix}(G_{1}) are nonempty and disjoint. Since XX is a complete CAT(0)\operatorname{CAT}(0) space, for i=0,1i=0,1 we have that BiB_{i} is a convex subset of XX [4, Corollary II.2.8(1)]. Note also that BiB_{i} is closed, hence complete in the induced metric, since GG acts isometrically. Hence Corollary 3.4 guarantees the existence of points a0B0a_{0}\in B_{0} and a1B1a_{1}\in B_{1} such that d(a0,a1)=d(B0,B1)>0d(a_{0},a_{1})=d(B_{0},B_{1})>0. Note that then a0a_{0} (respectively, a1a_{1}) is the closest-point projection of a1a_{1} (respectively, a0a_{0}) to B0B_{0} (respectively, B1B_{1}).

From now on, for i1i\geq 1 we write Δi\Delta_{i} for the spherical building which is the link of aia_{i} in XX, and xix_{i} for the point of Δi\Delta_{i} corresponding to the geodesic segment [ai1,ai][a_{i-1},a_{i}] (it will be seen from the construction that ai1a_{i-1} and aia_{i} are always distinct). By our assumption that every point of XX is a special vertex, each Δi\Delta_{i} will have the same type. Note that the Alexandrov angle at aia_{i} between any two points xx and yy of Δi\Delta_{i}, denoted ai(x,y)\angle_{a_{i}}(x,y), is equal to the distance between xx and yy in the CAT(1)\operatorname{CAT}(1) metric on Δi\Delta_{i}.

If Δ1\Delta_{1} is of type C2C_{2}, let p1p_{1} be a panel of Δ1\Delta_{1} at minimum distance from x1x_{1}. If Δ1\Delta_{1} is of type A2A_{2}, let c1c_{1} be a chamber of Δ1\Delta_{1} containing x1x_{1} and let p1p_{1} be a panel of c1c_{1}. The next, crucial, result uses the local lemmas from Section 2.

Lemma 3.11.
  1. (1)

    If Δ1\Delta_{1} is of type C2C_{2}, then there is a g1G1g_{1}\in G_{1} mapping p1p_{1} to a panel of Δ1\Delta_{1} which is opposite p1p_{1}.

  2. (2)

    If Δ1\Delta_{1} is of type A2A_{2}, then there is a g1G1g_{1}\in G_{1} mapping c1c_{1} to a chamber of Δ1\Delta_{1} which contains a panel p1opp_{1}^{\mathrm{op}} opposite p1p_{1}.

Moreover, in both cases, a1(p1,g1x1)α\angle_{a_{1}}(p_{1},g_{1}x_{1})\geq\alpha.

Proof.

Since G1G_{1} fixes a1a_{1}, it acts on Δ1\Delta_{1}. It suffices to show that case (2) in Lemma 2.1 or Lemma 2.2 does not occur (with the obvious modifications of notation). Assume by contradiction that there is a panel pp^{\prime} of Δ1\Delta_{1} which is fixed by G1G_{1} and is such that d(p,x1)<π/2d(p^{\prime},x_{1})<\pi/2 in the CAT(1)\operatorname{CAT}(1) metric on Δ1\Delta_{1}.

Since pp^{\prime} is fixed, every generator of G1G_{1} will fix a line segment [a1,xi][a_{1},x_{i}] with xia1x_{i}\neq a_{1}, and since G1G_{1} is finitely generated there exists a jj such that [a1,xj]B1[a_{1},x_{j}]\subset B_{1}. Hence there exists an x:=xja1B1x^{\prime}:=x_{j}\neq a_{1}\in B_{1} and we have a1(a0,x)<π/2\angle_{a_{1}}(a_{0},x^{\prime})<\pi/2. This contradicts  [4, Proposition II.2.4(3)], completing the proof. ∎

Let g1G1g_{1}\in G_{1} be as given by Lemma 3.11, and define a2:=g1a0a_{2}:=g_{1}a_{0}.

It will be helpful from here on to abuse terminology, as follows. If pp is a panel of Δi\Delta_{i} and MM is the wall of an apartment of XX containing aia_{i} which is determined by pp, then we will say that MM contains pp. We similarly abuse terminology for geodesic rays based at aia_{i} which are contained in walls determined by pp.

We now construct ξ\xi, and prove Lemma 3.10 for i=1i=1. Let S12S_{12} be a sector of XX based at a1a_{1} which contains the point a2a_{2}, so that in type A~2\tilde{A}_{2}, if a2a_{2} lies on a wall through a1a_{1} then S12S_{12} is a sector corresponding to the chamber g1c1g_{1}c_{1} of Δ1\Delta_{1}.

Lemma 3.12.

There is an apartment 𝒜1\mathcal{A}_{1} containing S12S_{12}, such that p1p_{1} is contained in a wall M1M_{1} of 𝒜1\mathcal{A}_{1}.

Proof.

By Lemma 3.11 we have that a1(p1,g1x1)α\angle_{a_{1}}(p_{1},g_{1}x_{1})\geq\alpha, hence the sectors S12S_{12} and g11(S12)g_{1}^{-1}(S_{12}) are distinct. So by [21, Proposition 1.15], there is an apartment 𝒜1\mathcal{A}_{1} containing both S12S_{12} and a germ of g11(S12)g_{1}^{-1}(S_{12}). Now define M1M_{1} to be the wall of 𝒜1\mathcal{A}_{1} through a1a_{1} which contains p1p_{1}. ∎

We define ξX\xi\in\partial X to be the endpoint of M1M_{1} such that the ray [a1,ξ)[a_{1},\xi) contains p1p_{1}. By construction a1(ξ,a2)=a1(p1,g1x1)\angle_{a_{1}}(\xi,a_{2})=\angle_{a_{1}}(p_{1},g_{1}x_{1}). Hence by Lemma 3.11 we have a1(ξ,a2)α\angle_{a_{1}}(\xi,a_{2})\geq\alpha. This proves Lemma 3.10 for i=1i=1.

For i1i\geq 1, we now define 𝐫i\mathbf{r}_{i} to be the ray [ai,ξ)[a_{i},\xi). By construction, 𝐫1=[a1,ξ)\mathbf{r}_{1}=[a_{1},\xi) is contained in a wall, and so by Lemma 3.9 and induction, for all i1i\geq 1 the ray 𝐫i\mathbf{r}_{i} will be contained in a wall. We may thus, for i1i\geq 1, define pip_{i} to be the panel of Δi\Delta_{i} induced by 𝐫i\mathbf{r}_{i}. By Lemma 3.9 again, if XX is of type C~2\tilde{C}_{2}, the panels pip_{i} are all of the same type.

We next prove Lemma 3.10 for i=2i=2. Define M2M_{2}^{\prime} to be the wall of 𝒜1\mathcal{A}_{1} containing 𝐫2\mathbf{r}_{2}.

Lemma 3.13.
  1. (1)

    If Δ2\Delta_{2} is of type C2C_{2}, then p2p_{2} is a panel of Δ2\Delta_{2} which is at minimum distance from x2x_{2}.

  2. (2)

    If Δ2\Delta_{2} is of type A2A_{2}, then there is a chamber c2c_{2} of Δ2\Delta_{2} which contains x2x_{2} such that p2p_{2} is a panel of c2c_{2}.

Proof.

If Δ2\Delta_{2} is of type C2C_{2} (see Figure 2), then by construction the wall M1M_{1} through a1a_{1} contains both p1p_{1} and g1p1g_{1}p_{1}. Now a1(x1,p1)=a1(g1x1,g1p1)\angle_{a_{1}}(x_{1},p_{1})=\angle_{a_{1}}(g_{1}x_{1},g_{1}p_{1}), both g1x1g_{1}x_{1} and x2x_{2} lie on the geodesic segment [a1,a2][a_{1},a_{2}], and M1M_{1} and M2M_{2}^{\prime} are parallel. It follows that a2(x2,p2)=a1(x1,p1)\angle_{a_{2}}(x_{2},p_{2})=\angle_{a_{1}}(x_{1},p_{1}), and thus p2p_{2} is a panel of Δ2\Delta_{2} at minimum distance from x2x_{2}.


\begin{overpic}[width=260.17464pt]{Lemma312typeC2} \put(-3.0,0.0){$\xi$} \put(-3.0,13.0){$\xi$} \put(10.0,-3.0){$M_{2}^{\prime}$} \put(10.0,16.0){$M_{1}$} \put(30.0,-2.0){$\mathbf{r}_{2}$} \put(30.0,16.0){$\mathbf{r}_{1}$} \put(85.0,-3.0){$p_{2}$} \put(62.0,17.0){$p_{1}$} \put(98.0,-2.0){$a_{2}$} \put(72.0,16.0){$a_{1}$} \put(90.0,7.0){$x_{2}$} \put(80.0,16.0){$g_{1}p_{1}$} \end{overpic}
Figure 2. The proof of Lemma 3.13 in type C2C_{2}.
\begin{overpic}[width=260.17464pt]{Lemma312typeA2} \put(-3.0,0.0){$\xi$} \put(-3.0,16.0){$\xi$} \put(10.0,-3.0){$M_{2}^{\prime}$} \put(10.0,19.0){$M_{1}$} \put(30.0,-2.0){$\mathbf{r}_{2}$} \put(30.0,19.0){$\mathbf{r}_{1}$} \put(85.0,-3.0){$p_{2}$} \put(98.0,-2.0){$a_{2}$} \put(84.0,19.0){$a_{1}$} \put(91.0,5.0){$c_{2}$} \put(55.0,25.0){$S_{2}$} \end{overpic}
Figure 3. The proof of Lemma 3.13 in type A2A_{2}.

If Δ2\Delta_{2} is of type A2A_{2} (see Figure 3), there is a sector S2S_{2} of 𝒜1\mathcal{A}_{1} which is based at a2a_{2} and is bounded by 𝐫2\mathbf{r}_{2}, such that a1a_{1} is in S2S_{2}. Then we take c2c_{2} to be the chamber of Δ2\Delta_{2} determined by S2S_{2}. ∎

Now define G2:=g1G0g11G_{2}:=g_{1}G_{0}g_{1}^{-1} and B2:=g1B0B_{2}:=g_{1}B_{0}, so that B2B_{2} is the fixed set of G2G_{2}. Then since g1g_{1} is an isometry which fixes B1B_{1}, we have d(B0,B1)=d(B1,B2)=d(a1,a2)d(B_{0},B_{1})=d(B_{1},B_{2})=d(a_{1},a_{2}), and that a2a_{2} is the closest-point projection of a1a_{1} to B2B_{2}. The next result is proved using the local lemmas of Section 2 and similar arguments to those in the proof of Lemma 3.11.

Lemma 3.14.
  1. (1)

    If Δ2\Delta_{2} is of type C2C_{2}, there is a g2G2g_{2}\in G_{2} mapping p2p_{2} to a panel of Δ2\Delta_{2} which is opposite p2p_{2}.

  2. (2)

    If Δ2\Delta_{2} is of type A2A_{2}, let c2c_{2} be as given by Lemma 3.13(2). Then there is a g2G2g_{2}\in G_{2} mapping c2c_{2} to a chamber of Δ2\Delta_{2} which contains a panel p2opp_{2}^{\mathrm{op}} opposite p2p_{2}.

Moreover, in all cases, a2(p2,g2x2)α\angle_{a_{2}}(p_{2},g_{2}x_{2})\geq\alpha.

We now define

g:=g2g1G.g:=g_{2}g_{1}\in G.

Observe that since g1g_{1} fixes a1a_{1} and g2g_{2} fixes a2=g1a0a_{2}=g_{1}a_{0}, we have a2=ga0a_{2}=ga_{0} and a3=ga1a_{3}=ga_{1}. We may thus, for i2i\geq 2, inductively define

ai:=gai2.a_{i}:=ga_{i-2}.

We also for i3i\geq 3 define

gi:=ggi2g1.g_{i}:=gg_{i-2}g^{-1}.

An easy induction shows that gi1gi2=gg_{i-1}g_{i-2}=g for all i3i\geq 3. Also, for all i1i\geq 1, by induction gig_{i} fixes aia_{i}, hence gig_{i} acts on Δi\Delta_{i}, and we have giai1=ai+1g_{i}a_{i-1}=a_{i+1}. Finally, if Δi\Delta_{i} is of type A2A_{2} then for i3i\geq 3 we define (with abuse of terminology using chambers in the spherical buildings instead of their corresponding sectors in XX)

ci:=gci2.c_{i}:=gc_{i-2}.

The next result completes the proof of Lemma 3.10.

Lemma 3.15.

For i1i\geq 1:

  1. (1)

    If Δi\Delta_{i} is of type C2C_{2}, then pip_{i} is equal to gpi2gp_{i-2} and is a panel of Δi\Delta_{i} which is at minimum distance from xix_{i}. Moreover, gipig_{i}p_{i} is opposite pip_{i}.

  2. (2)

    If Δi\Delta_{i} is of type A2A_{2}, then cic_{i} contains xix_{i}, pip_{i} is a panel of cic_{i} and gicig_{i}c_{i} contains a panel piopp_{i}^{\mathrm{op}} which is opposite pip_{i}.

Moreover, in all cases, ai(ξ,ai+1)α\angle_{a_{i}}(\xi,a_{i+1})\geq\alpha.

Proof.

The proof is by induction on ii, and the cases i=1,2i=1,2 have been established above.


\begin{overpic}[width=260.17464pt]{Lemma315typeC2v2} \put(-3.0,0.0){$\xi$} \put(-3.0,13.0){$\xi$} \put(10.0,-3.0){$M_{i}^{\prime}$} \put(10.0,16.0){$M_{i-1}$} \put(30.0,-2.0){$\mathbf{r}_{i}$} \put(30.0,16.0){$\mathbf{r}_{i-1}$} \put(85.0,-3.0){$p_{i}$} \put(62.0,18.0){$p_{i-1}$} \put(98.0,-2.0){$a_{i}$} \put(72.0,16.0){$a_{i-1}$} \put(90.0,7.0){$x_{i}$} \end{overpic}
Figure 4. The proof of Lemma 3.15 in type C2C_{2}.

Suppose first that Δi\Delta_{i} is of type C2C_{2}. For i3i\geq 3, by induction the panel gpi2gp_{i-2} is at minimum distance from gxi2=xigx_{i-2}=x_{i}. Since the action of GG is type-preserving, gpi2gp_{i-2} has the same type as pi2p_{i-2}. Thus pip_{i} and gpi2gp_{i-2} have the same type.

We next show that pip_{i} is a panel at minimum distance from xix_{i} (see Figure 4). Let Si1S_{i-1} be a sector of XX based at ai1a_{i-1} which contains the point aia_{i}. By induction, the panels pi1p_{i-1} and gi1pi1g_{i-1}p_{i-1} are opposite, and so the union of 𝐫i1\mathbf{r}_{i-1} with the facet of Si1S_{i-1} containing gi1pi1g_{i-1}p_{i-1} is a wall Mi1M_{i-1} of XX. Moreover the wall Mi1M_{i-1} bounds both Si1S_{i-1} and a sector Si1opS_{i-1}^{\mathrm{op}} of XX which is opposite Si1S_{i-1}. Thus by [21, Proposition 1.12], there exists a (unique) apartment 𝒜i1\mathcal{A}_{i-1} containing both Si1opS_{i-1}^{\mathrm{op}}, hence 𝐫i1\mathbf{r}_{i-1}, and Si1S_{i-1}. Let MiM_{i}^{\prime} be the wall of 𝒜i1\mathcal{A}_{i-1} containing 𝐫i\mathbf{r}_{i}. Now observe that as Mi1M_{i-1} and MiM_{i}^{\prime} are parallel, we have ai1(gi1pi1,gi1xi1)=ai(pi,xi)\angle_{a_{i-1}}(g_{i-1}p_{i-1},g_{i-1}x_{i-1})=\angle_{a_{i}}(p_{i},x_{i}). Also, by induction, gi1pi1g_{i-1}p_{i-1} is a panel at minimum distance from gi1xi1g_{i-1}x_{i-1}. Therefore pip_{i} is a panel at minimal distance from xix_{i}.

Now pip_{i} and gpi2gp_{i-2} are panels of the same type both at minimum distance from xix_{i}. It follows that pi=gpi2p_{i}=gp_{i-2}. Hence as pi2p_{i-2} and gi2pi2g_{i-2}p_{i-2} are opposite, the panels pi=gpi2p_{i}=gp_{i-2} and ggi2pi2gg_{i-2}p_{i-2} are opposite. To complete the proof of (1) in type C2C_{2}, we observe that

gipi=ggi2g1gpi2=ggi2pi2.g_{i}p_{i}=gg_{i-2}g^{-1}gp_{i-2}=gg_{i-2}p_{i-2}.

If Δi\Delta_{i} is of type A2A_{2}, then since ci=gci2c_{i}=gc_{i-2} and xi=gxi2x_{i}=gx_{i-2}, by induction cic_{i} contains xix_{i}. We next show that pi=gpi2p_{i}=gp_{i-2}, which implies that pip_{i} is a panel of cic_{i}. Now ci1c_{i-1} is a chamber of Δi1\Delta_{i-1} which contains xi1x_{i-1} and pi1p_{i-1} is a panel of ci1c_{i-1}. Hence as GG is type-preserving and opposite panels in type A2A_{2} have distinct types, gi1ci1g_{i-1}c_{i-1} has panels pi1opp_{i-1}^{\mathrm{op}} and gi1pi1g_{i-1}p_{i-1} (and contains gi1xi1g_{i-1}x_{i-1}). Let Si1S_{i-1} be a sector of XX which is based at ai1a_{i-1} and contains ai2a_{i-2}. We consider two cases.


\begin{overpic}[width=260.17464pt]{Lemma315typeA2opposite} \put(-3.0,44.0){$\xi$} \put(-3.0,33.0){$\xi$} \put(-3.0,22.0){$\xi$} \put(8.0,44.0){\footnotesize{$M_{i}^{\prime\prime}$}} \put(8.0,36.0){\footnotesize{$M_{i-1}^{\prime\prime}$}} \put(8.0,25.0){\footnotesize{$M_{i-2}^{\prime\prime}$}} \put(69.0,45.0){\footnotesize{$p_{i}$}} \put(51.0,36.0){\footnotesize{$p_{i-1}$}} \put(35.0,26.0){\footnotesize{$p_{i-2}$}} \put(34.0,31.0){\footnotesize{$g_{i-2}p_{i-2}$}} \put(79.0,43.0){\footnotesize{$a_{i}$}} \put(63.0,32.0){\footnotesize{$a_{i-1}$}} \put(46.0,23.0){\footnotesize{$a_{i-2}$}} \put(35.0,12.0){\footnotesize{$S_{i-1}$}} \put(76.0,54.0){\footnotesize{$g_{i-1}S_{i-1}$}} \put(43.0,68.0){\footnotesize{$r_{i-1}$}} \end{overpic}
Figure 5. The proof of Lemma 3.15 in type A2A_{2}, Case I.

Case I: gi1ci1g_{i-1}c_{i-1} and ci1c_{i-1} are opposite. See Figure 5. In this case, since gi1Si1g_{i-1}S_{i-1} and Si1S_{i-1} are opposite, by [21, Proposition 1.12] there is a unique apartment 𝒜i1\mathcal{A}_{i-1} which contains both Si1S_{i-1} and gi1Si1g_{i-1}S_{i-1}. Let ri1r_{i-1} be the reflection of 𝒜i1\mathcal{A}_{i-1} in its unique wall which passes through ai1a_{i-1} and does not bound Si1S_{i-1} (or gi1Si1g_{i-1}S_{i-1}). That is, ri1r_{i-1} is the reflection of 𝒜i1\mathcal{A}_{i-1} which fixes ai1a_{i-1} and takes pi1p_{i-1} to gi1pi1g_{i-1}p_{i-1}. Then the geodesic segment [ai1,ai]=gi1[ai1,ai2][a_{i-1},a_{i}]=g_{i-1}[a_{i-1},a_{i-2}] of 𝒜i1\mathcal{A}_{i-1} is obtained from [ai1,ai2][a_{i-1},a_{i-2}] by applying the reflection ri1r_{i-1}. For j=i2,i1,ij=i-2,i-1,i, write Mj′′M_{j}^{\prime\prime} for the wall of 𝒜i1\mathcal{A}_{i-1} which passes through aja_{j} and contains pjp_{j}. Then since each Mj′′M_{j}^{\prime\prime} contains at least some initial portion of the geodesic ray 𝐫j\mathbf{r}_{j}, the three walls Mi2′′M_{i-2}^{\prime\prime}, Mi1′′M_{i-1}^{\prime\prime} and Mi′′M_{i}^{\prime\prime} of 𝒜i1\mathcal{A}_{i-1}^{\prime} are mutually parallel. It follows that ri1r_{i-1} maps the wall of 𝒜i1\mathcal{A}_{i-1} which passes through ai2a_{i-2} and contains gi2pi2g_{i-2}p_{i-2} to Mi′′M_{i}^{\prime\prime}. Hence gi1gi2pi2=gpi2=pig_{i-1}g_{i-2}p_{i-2}=gp_{i-2}=p_{i} in this case.


\begin{overpic}[width=260.17464pt]{Lemma315typeA2nonopposite} \put(-3.0,33.0){$\xi$} \put(-3.0,22.0){$\xi$} \put(51.0,36.0){\footnotesize{$p_{i-1}$}} \put(35.0,26.0){\footnotesize{$p_{i-2}$}} \put(34.0,31.0){\footnotesize{$g_{i-2}p_{i-2}$}} \put(78.0,26.0){\footnotesize{$\rho_{i-1}(a_{i})$}} \put(64.0,22.5){\footnotesize{$\rho_{i-1}(p_{i})$}} \put(60.0,35.0){\footnotesize{$a_{i-1}$}} \put(46.0,23.0){\footnotesize{$a_{i-2}$}} \put(35.0,12.0){\footnotesize{$S_{i-1}$}} \put(78.0,15.0){\footnotesize{$\rho_{i-1}(g_{i-1}S_{i-1})$}} \end{overpic}
Figure 6. The proof of Lemma 3.15 in type A2A_{2}, Case II.

Case II: gi1ci1g_{i-1}c_{i-1} and ci1c_{i-1} are not opposite. See Figure 6. Note that these chambers cannot be adjacent in Δi1\Delta_{i-1}, as gi1ci1g_{i-1}c_{i-1} contains a panel opposite to the panel pi1p_{i-1} of ci1c_{i-1}. By [21, Proposition 1.15], there is an apartment 𝒜i1\mathcal{A}_{i-1} which contains Si1S_{i-1} and a germ of gi1Si1g_{i-1}S_{i-1}. Write ρi1\rho_{i-1} for the retraction of XX onto 𝒜i1\mathcal{A}_{i-1} such that ρi11(ai1)={ai1}\rho_{i-1}^{-1}(a_{i-1})=\{a_{i-1}\}, as guaranteed by [21, Axiom (A5)]. Then by [21, Proposition 1.17], ρi1\rho_{i-1} maps gi1Si1g_{i-1}S_{i-1} isometrically onto the sector ρi1(gi1Si1)\rho_{i-1}(g_{i-1}S_{i-1}) of 𝒜i1\mathcal{A}_{i-1} which is based at ai1a_{i-1} and has the same germ as gi1Si1g_{i-1}S_{i-1}. Thus d(ai1,ai2)=d(ai1,ai)=d(ai1,ρi1(ai))d(a_{i-1},a_{i-2})=d(a_{i-1},a_{i})=d(a_{i-1},\rho_{i-1}(a_{i})), and Si1S_{i-1} and ρi1(gi1Si1)\rho_{i-1}(g_{i-1}S_{i-1}) are nonadjacent and nonopposite sectors of 𝒜i1\mathcal{A}_{i-1} which respectively contain the geodesic segments [ai1,ai2][a_{i-1},a_{i-2}] and [ai1,ρi1(ai)][a_{i-1},\rho_{i-1}(a_{i})]. Hence [ai1,ρi1(ai)][a_{i-1},\rho_{i-1}(a_{i})] is obtained from [ai1,ai2][a_{i-1},a_{i-2}] by applying a rotation of 𝒜i1\mathcal{A}_{i-1} about the point ai1a_{i-1} through angle 2π/32\pi/3. This rotation takes the wall of 𝒜i1\mathcal{A}_{i-1} through ai2a_{i-2} which contains gi2pi2g_{i-2}p_{i-2} to the wall of 𝒜i1\mathcal{A}_{i-1} through ρi1(ai)\rho_{i-1}(a_{i}) which contains ρi1(pi)\rho_{i-1}(p_{i}). Since retractions and this rotation are type-preserving, it follows that gi1gi2pi2=gpi2=pig_{i-1}g_{i-2}p_{i-2}=gp_{i-2}=p_{i}, as required.

We have now shown that pi=gpi2p_{i}=gp_{i-2}. Since gi2ci2g_{i-2}c_{i-2} contains pi2opp_{i-2}^{\mathrm{op}}, which is opposite pi2p_{i-2}, we have that ggi2ci2gg_{i-2}c_{i-2} contains g(pi2op)g(p_{i-2}^{\mathrm{op}}), which is opposite pi=gpi2p_{i}=gp_{i-2}. Also,

gici=ggi2g1gci2=ggi2ci2.g_{i}c_{i}=gg_{i-2}g^{-1}gc_{i-2}=gg_{i-2}c_{i-2}.

So gicig_{i}c_{i} contains a panel piopp_{i}^{\mathrm{op}} which is opposite pip_{i}, as required to finish the proof of (2).

Now the same arguments as in the cases i=1,2i=1,2 show that ai(pi,gixi)α\angle_{a_{i}}(p_{i},g_{i}x_{i})\geq\alpha. By construction we have ai(pi,gixi)=ai(ξ,ai+1)\angle_{a_{i}}(p_{i},g_{i}x_{i})=\angle_{a_{i}}(\xi,a_{i+1}), which completes the proof. ∎

Remark 3.16.

Despite substantial effort we have been unable to extend our approach to type G~2\tilde{G}_{2}. First, as explained in Remark 2.3, there cannot be a local lemma for G2G_{2} which guarantees that we obtain a panel opposite to pip_{i} at every step. Hence we cannot make use of opposite sectors in XX as we did in type C~2\tilde{C}_{2} above.

Moreover, the local lemma that we have been able to prove in type G2G_{2} (see Lemma 2.4) does not give us enough control to be able to show that pi=gpi2p_{i}=gp_{i-2}, as we did in type A~2\tilde{A}_{2}. The real issue is that for XX of type G~2\tilde{G}_{2}, if possibility (2) in Lemma 2.4 occurs at step ii, then the rays 𝐫i\mathbf{r}_{i} and 𝐫i+1\mathbf{r}_{i+1} have germs which may or may not be parallel in an apartment of XX which contains both of these germs. This makes it very difficult to run an inductive procedure. We have also tried defining g=g4g3g2g1g=g_{4}g_{3}g_{2}g_{1} in various ways, rather than using g=g2g1g=g_{2}g_{1} as above, but this just postpones the problem.

3.4. Unbounded Busemann function

For any metric space YY, the Busemann function (see [4, Definition II.8.17]) associated to a geodesic ray γ\gamma in YY is given by, for yYy\in Y,

bγ(y):=limt[d(y,γ(t))t].b_{\gamma}(y):=\lim_{t\to\infty}[d(y,\gamma(t))-t].

We will apply the following general result.

Lemma 3.17.

Let YY be a complete CAT(0)\operatorname{CAT}(0) space. Let yYy\in Y, let ηY\eta\in\partial Y and let γ\gamma be the geodesic ray [y,η)[y,\eta). Suppose zYz\in Y is such that d(y,z)=D>0d(y,z)=D>0 and y(η,z)=θ>π2\angle_{y}(\eta,z)=\theta>\frac{\pi}{2}. Then

bγ(z)Dsin(θπ2)>0.b_{\gamma}(z)\geq D\sin\left(\theta-\frac{\pi}{2}\right)>0.
Proof.

Fix t>0t>0 and write yt=γ(t)y_{t}=\gamma(t), so that θ=y(η,z)=y(yt,z)\theta=\angle_{y}(\eta,z)=\angle_{y}(y_{t},z). Consider a triangle in the Euclidean plane 𝔼2\mathbb{E}^{2} with vertices yt^\hat{y_{t}}, y^\hat{y} and z^\hat{z}, so that d(y^,yt^)=d(y,yt)=td(\hat{y},\hat{y_{t}})=d(y,y_{t})=t, d(y^,z^)=d(y,z)=Dd(\hat{y},\hat{z})=d(y,z)=D and y^(yt^,z^)=θ\angle_{\hat{y}}(\hat{y_{t}},\hat{z})=\theta. Then by [4, Proposition II.1.7(5)], we have d(z,yt)d(z^,yt^)d(z,y_{t})\geq d(\hat{z},\hat{y_{t}}).

Let \ell be the line in 𝔼2\mathbb{E}^{2} which extends [yt^,y^][\hat{y_{t}},\hat{y}] and let p:𝔼2p_{\ell}:\mathbb{E}^{2}\to\ell be the closest-point projection. Then since θ>π2\theta>\frac{\pi}{2}, the point y^\hat{y} lies strictly between the points yt^\hat{y_{t}} and p(z^)p_{\ell}(\hat{z}) on \ell. If θ=π\theta=\pi, equivalently z^\hat{z} lies on \ell, then

d(z^,yt^)=d(z^,y^)+d(y^,yt^)=D+t=Dsin(θπ2)+t.d(\hat{z},\hat{y_{t}})=d(\hat{z},\hat{y})+d(\hat{y},\hat{y_{t}})=D+t=D\sin\left(\theta-\frac{\pi}{2}\right)+t.

Otherwise, by considering the right-angled Euclidean triangle with vertices y^\hat{y}, p(z^)p_{\ell}(\hat{z}) and z^\hat{z}, we calculate d(y^,p(z^))=Dsin(θπ2)d(\hat{y},p_{\ell}(\hat{z}))=D\sin(\theta-\frac{\pi}{2}). Hence as [z^,yt^][\hat{z},\hat{y_{t}}] is the hypotenuse of the right-angled Euclidean triangle with vertices yt^\hat{y_{t}}, p(z^)p_{\ell}(\hat{z}) and z^\hat{z}, we obtain

d(z^,yt^)>d(p(z^),yt^)=d(p(z^),y^)+d(y^,yt^)=Dsin(θπ2)+t.d(\hat{z},\hat{y_{t}})>d(p_{\ell}(\hat{z}),\hat{y_{t}})=d(p_{\ell}(\hat{z}),\hat{y})+d(\hat{y},\hat{y_{t}})=D\sin\left(\theta-\frac{\pi}{2}\right)+t.

We have now shown that for all t>0t>0,

d(z,γ(t))t=d(z,yt)tDsin(θπ2).d(z,\gamma(t))-t=d(z,y_{t})-t\geq D\sin\left(\theta-\frac{\pi}{2}\right).

Taking the limit as tt\to\infty gives the desired result. ∎

Now let XX be as in Section 3.3, and let {ai}i=0X\{a_{i}\}_{i=0}^{\infty}\subset X and ξX\xi\in\partial X be as constructed there. Recall that 𝐫i\mathbf{r}_{i} is the geodesic ray [ai,ξ)[a_{i},\xi).

Corollary 3.18.

limib𝐫𝟏(ai)=+\lim_{i\to\infty}b_{\mathbf{r_{1}}}(a_{i})=+\infty.

Proof.

Let D=d(a0,a1)=d(ai,ai+1)>0D=d(a_{0},a_{1})=d(a_{i},a_{i+1})>0. By Lemma 3.10, for all i1i\geq 1 we have ai(ξ,ai+1)α>π2\angle_{a_{i}}(\xi,a_{i+1})\geq\alpha>\frac{\pi}{2}. Thus by Lemma 3.17, for all i1i\geq 1

b𝐫i(ai+1)Dsin(απ2)>0.b_{\mathbf{r}_{i}}(a_{i+1})\geq D\sin\left(\alpha-\frac{\pi}{2}\right)>0.

As b𝐫i(ai)=0b_{\mathbf{r}_{i}}(a_{i})=0, in fact

b𝐫i(ai+1)b𝐫i(ai)Dsin(απ2)>0.b_{\mathbf{r}_{i}}(a_{i+1})-b_{\mathbf{r}_{i}}(a_{i})\geq D\sin\left(\alpha-\frac{\pi}{2}\right)>0.

Now for all i1i\geq 1 the rays 𝐫i\mathbf{r}_{i} are asymptotic, since they all have endpoint ξ\xi. Hence the Busemann functions b𝐫ib_{\mathbf{r}_{i}} pairwise differ by a constant [4, Corollary II.8.20]. Thus the difference b𝐫1(ai+1)b𝐫1(ai)b_{\mathbf{r}_{1}}(a_{i+1})-b_{\mathbf{r}_{1}}(a_{i}) has a strictly positive lower bound which is independent of ii, proving the result. ∎

3.5. End of proof of Proposition 3.8

If gg were elliptic then a0a_{0} would have a bounded orbit under gg. By definition, a2ka_{2k} equals gka0g^{k}a_{0} for all k0k\geq 0, and by Corollary 3.18 we have limkb𝐫1(a2k)=+\lim_{k\to\infty}b_{\mathbf{r}_{1}}(a_{2k})=+\infty. But b𝐫1(a2k)b𝐫1(a0)b_{\mathbf{r}_{1}}(a_{2k})-b_{\mathbf{r}_{1}}(a_{0}) is at most d(a0,a2k)d(a_{0},a_{2k}), a contradiction. Thus by [20, Corollaire 4.2], gg is hyperbolic.

Bibliography

  • [1] P. Abramenko and K. S. Brown, Buildings, vol. 248 of Graduate Texts in Mathematics, Springer, New York, 2008. Theory and applications.
  • [2] W. Ballmann and M. Brin, Orbihedra of nonpositive curvature, Inst. Hautes Études Sci. Publ. Math., (1995), pp. 169–209 (1996).
  • [3] E. Breuillard and K. Fujiwara, On the joint spectral radius for isometries of non-positively curved spaces and uniform growth, arXiv e-prints, (2018).
  • [4] M. R. Bridson and A. Haefliger, Metric spaces of non-positive curvature, vol. 319 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], Springer-Verlag, Berlin, 1999.
  • [5] F. Bruhat and J. Tits, Groupes réductifs sur un corps local : I. données radicielles valuées, Publications Mathématiques de l’IHÉS, 41 (1972), pp. 5–251.
  • [6] P.-E. Caprace and A. Lytchak, At infinity of finite-dimensional CAT(0) spaces, Math. Ann., 346 (2010), pp. 1–21.
  • [7] M. Culler and J. Morgan, Group actions on \mathbb{R}-trees, Proc. London. Math. Soc., 55 ((1987)), pp. 571–604.
  • [8] S. Ivanov, On Helly’s theorem in geodesic spaces, Electron. Res. Announc. Math. Sci., 21 (2014), pp. 109–112.
  • [9] B. Kleiner and B. Leeb, Rigidity of quasi-isometries for symmetric spaces and euclidean buildings, Publications Mathématiques de l’IHÉS, 86 (1997), pp. 115–197.
  • [10] L. Kramer, Metric properties of Euclidean buildings, in Global differential geometry, vol. 17 of Springer Proc. Math., Springer, Heidelberg, 2012, pp. 147–159.
  • [11] N. Leder and O. Varghese, A note on locally elliptic actions on cube complexes, Innov. Incidence Geom., 18 (2020), pp. 1–6.
  • [12] H. Leung, J. Schillewaert, and A. Thomas, Distances between fixed-point sets in 2-dimensional euclidean buildings are realised, (2022).
  • [13] T. Marquis, A fixed point theorem for Lie groups acting on buildings and applications to Kac-Moody theory, Forum Math., 27 (2015), pp. 449–466.
  • [14] B. Martin, J. Schillewaert, G. F. Steinke, and K. Struyve, On metrically complete Bruhat-Tits buildings, Adv. Geom., 13 (2013), pp. 497–510.
  • [15] H. Masur, Closed trajectories for quadratic differentials with an application to billiards, Duke Math. J., 53 (1986), pp. 307–314.
  • [16] J. W. Morgan and P. B. Shalen, Valuations, trees, and degenerations of hyperbolic structures. I, Ann. of Math. (2), 120 (1984), pp. 401–476.
  • [17] B. Mühlherr, K. Struyve, and H. Van Maldeghem, Descent of affine buildings - i. large minimal angles, Trans. Amer. Math. Soc., 366 (2014), pp. 4345–4366.
  • [18] S. Norin, D. Osajda, and P. Przytycki, Torsion groups do not act on 2-dimensional CAT(0) complexes, Duke Math. J., 171 (2022), pp. 1379–1415.
  • [19] D. Osajda, Group cubization, Duke Math. J., 167 (2018), pp. 1049–1055. With an appendix by Mikaël Pichot.
  • [20] A. Parreau, Immeubles affines: construction par les normes et étude des isométries, in Crystallographic groups and their generalizations (Kortrijk, 1999), vol. 262 of Contemp. Math., Amer. Math. Soc., Providence, RI, 2000, pp. 263–302.
  • [21]  , Sous-groupes elliptiques de groupes linéaires sur un corps valué, J. Lie Theory, 13 (2003), pp. 271–278.
  • [22]  , Affine buildings: construction by norms and study of isometries, (2022). Translated from the French by Harris Leung, Jeroen Schillewaert and Anne Thomas.
  • [23] M. Ronan, Lectures on buildings, University of Chicago Press, Chicago, IL, 2009. Updated and revised.
  • [24] G. Rousseau, Euclidean buildings, in Géométries à courbure négative ou nulle, groupes discrets et rigidités, vol. 18 of Sémin. Congr., Soc. Math. France, Paris, 2009, pp. 77–116.
  • [25] M. Sageev, Ends of group pairs and non-positively curved cube complexes, Proc. London Math. Soc. (3), 71 (1995), pp. 585–617.
  • [26] J.-P. Serre, Trees, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2003. Translated from the French original by John Stillwell, Corrected 2nd printing of the 1980 English translation.
  • [27] K. Struyve, (Non)-completeness of \mathbb{R}-buildings and fixed point theorems, Groups Geom. Dyn., 5 (2011), pp. 177–188.
  • [28] E. L. Swenson, A cut point theorem for CAT(0)\rm{CAT}(0) groups, J. Differential Geom., 53 (1999), pp. 327–358.
  • [29] J. Tits, Immeubles de type affine, in Buildings and the Geometry of Diagrams, Springer Lecture Notes, vol. 1181, Springer Verlag, 1986, pp. 159–190.