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

Lipschitz connectivity and filling invariants in solvable groups and buildings

Robert Young
(Date: September 30, 2025)
Abstract.

We give some new methods, based on Lipschitz extension theorems, for bounding filling invariants of subsets of nonpositively curved spaces. We apply our methods to find sharp bounds on higher-order Dehn functions of Sol2n+1\operatorname{Sol}_{2n+1}, horospheres in euclidean buildings, Hilbert modular groups, and certain SS-arithmetic groups.

1. Introduction

Filling invariants of a group or space, such as Dehn functions and higher-order Dehn functions, are quantitative versions of finiteness properties. There are many methods for bounding the Dehn function, but bounds on the Dehn function are often difficult to generalize to higher-order Dehn functions. For example, one can prove that a non-positively curved space has a Dehn function which is at most quadratic in a couple of lines: the fact that the distance function is convex implies that the disc formed by connecting every point on the curve to a basepoint on the curve has quadratically large area. On the other hand, proving that a non-positively curved space has a kkth-order Dehn function bounded by V(k+1)/kV^{(k+1)/k} takes several pages [Gro83, Wen08]. In this paper, we describe some new methods for bounding higher-order Dehn functions and apply them to solvable groups and subsets of nonpositively curved spaces.

One reason that higher-order Dehn functions are harder to bound is that the geometry of spheres is more complicated than the geometry of curves. A closed curve is geometrically very simple. It has diameter bounded by its length, it has a natural parameterization by length, and a closed curve in a space with a geometric group action can be approximated by a word in the group. None of these hold for spheres. A kk-sphere of volume VV may have arbitrarily large diameter, has no natural parameterization, and, though it can often be approximated by a cellular or simplicial sphere, that sphere may have arbitrarily many cells of dimension less than kk.

One way around this is to consider Lipschitz extension properties. A typical Lipschitz extension property is Lipschitz kk-connectivity; we say that a space XX is Lipschitz kk-connected (with constant cc) if there is a cc such that for any 0dk0\leq d\leq k and any ll-Lipschitz map f:SdXf:S^{d}\to X, there is a clcl-Lipschitz extension f¯:Dd+1X\bar{f}:D^{d+1}\to X. The advantage of dealing with Lipschitz spheres rather than spheres of bounded volume is that techniques for filling closed curves often generalize to Lipschitz spheres. For example, the same construction that shows that a non-positively curved space has quadratic Dehn function shows that such a space is Lipschitz kk-connected for any kk. Any map f:SdXf:S^{d}\to X can be extended to a map f¯:Dd+1X\bar{f}:D^{d+1}\to X by coning ff off to a point along geodesics, and if ff is Lipschitz, so is f¯\bar{f}.

In this paper, we describe a way to use Lipschitz connectivity to prove bounds on higher-order filling functions of subsets of spaces with finite Assouad-Nagata dimension. These spaces include euclidean buildings and homogeneous Hadamard manifolds [LS05], and we will show that a higher-dimensional analogue of the Lubotzky-Mozes-Raghunathan theorem holds for Lipschitz nn-connected subsets of spaces with finite Assouad-Nagata dimension. Recall that Lubotzky, Mozes, and Raghunathan proved that

Theorem 1.1.

[LMR00] If Γ\Gamma is an irreducible lattice in a semisimple group GG of rank 2\geq 2, then the word metric on Γ\Gamma is quasi-isometric to the restriction of the metric on GG to Γ\Gamma.

One way to state this theorem is that the inclusion ΓG\Gamma\hookrightarrow G does not induce any distortion of lengths. That is, there is a c>0c>0 such that if x,yΓx,y\in\Gamma are connected by a path of length ll in GG, then they are connected by a path of length cl\leq cl in the Cayley graph of Γ\Gamma. We can think of this as an efficient 1-dimensional filling of a 0-sphere. Many authors have conjectured that when GG has higher rank, we can fill higher-dimensional spheres efficiently; for example, Thurston famously conjectured that SL(n;)\operatorname{SL}(n;\mathbb{Z}) has quadratic Dehn function for n4n\geq 4 [Ger93], and Gromov conjectured that the (k2)(k-2)-th order Dehn function of a lattice in a symmetric space of rank kk should be bounded by a polynomial [Gro93]. Bux and Wortman [BW07] conjectured that filling volumes should be undistorted in lattices in higher-rank semisimple groups. We will state a version of this conjecture in terms of homological filling volumes; in a highly-connected space, these are equivalent to homotopical filling volumes in dimensions above 2 [Gro83, Whi84, Gro].

To state the conjecture, we introduce Lipschitz chains. A Lipschitz dd-chain in YY is a formal sum of Lipschitz maps ΔdY\Delta_{d}\to Y. One can define the boundary of a Lipschitz chain as for singular chains, and this gives rise to a homology theory. If α\alpha is a Lipschitz dd-cycle in YY, define

FVYd+1(α)=infβ=αmassβ.\operatorname{FV}^{d+1}_{Y}(\alpha)=\inf_{\partial\beta=\alpha}\operatorname{mass}\beta.

to be the filling volume of α\alpha in YY. In particular, if YY is a geodesic metric space and α\alpha is the 0-cycle xyx-y, then FVY1(α)=d(x,y)\operatorname{FV}^{1}_{Y}(\alpha)=d(x,y).

If ZXZ\subset X, we say that ZZ is undistorted up to dimension nn if there is some r0r\geq 0 and cc such that if α\alpha is a Lipschitz dd-cycle in ZZ and d<nd<n, then

FVZd+1(α)cFVXd+1(α)+c.\operatorname{FV}^{d+1}_{Z}(\alpha)\leq c\operatorname{FV}^{d+1}_{X}(\alpha)+c.

(Note that this differs from Bux and Wortman’s definition in [BW07]; Bux and Wortman’s definition deals with extending spheres in a neighborhood of ZZ to balls in a larger neighborhood.)

Conjecture 1.2 (see [BW07], Question 1.6).

If Γ\Gamma is an irreducible lattice in a semisimple group GG of rank nn, then there is a nonempty Γ\Gamma-invariant subset ZGZ\subset G such that dHaus(Z,Γ)<d_{\text{Haus}}(Z,\Gamma)<\infty and ZZ is undistorted up to dimension n1n-1.

Here, dHaus(Z,Γ)d_{\text{Haus}}(Z,\Gamma) represents the Hausdorff distance between the two sets.

Theorem 1.1 is a special case of this conjecture. As Bestvina, Eskin, and Wortman note in [BEW], Conj. 1.2 would imply that the kkth-order Dehn function of Γ\Gamma is bounded by V(k+1)/kV^{(k+1)/k}. In recent years, a significant amount of progess has been made toward these conjectures. Druţu proved that a lattice of \mathbb{Q}-rank 1 in a symmetric space of \mathbb{R}-rank 3\geq 3 has a Dehn function bounded by n2+ϵn^{2+\epsilon} for any ϵ>0\epsilon>0 [Dru04], Leuzinger and Pittet proved that, conversely, any irreducible lattice in a symmetric space of rank 2 which is not cocompact has an exponentially large Dehn function [LP96], and the author proved Thurston’s conjecture in the case that n5n\geq 5 [You].

In this paper, we make a step toward proving Conj. 1.2 by showing that, under some conditions on GG and Γ\Gamma, undistortedness follows from a Lipschitz extension property. We say that ZZ is Lipschitz nn-connected if there is a cc such that for any 0dk0\leq d\leq k and any ll-Lipschitz map f:SdZf:S^{d}\to Z, there is a clcl-Lipschitz extension f¯:Dd+1Z\bar{f}:D^{d+1}\to Z. If ZYZ\subset Y, we say ZZ is Lipschitz nn-connected in YY if, under the above conditions, there is a clcl-Lipschitz extension f¯:Dd+1Y\bar{f}:D^{d+1}\to Y.

Theorem 1.3.

Suppose that ZXZ\subset X is a nonempty closed subset with metric given by the restriction of the metric of XX. Suppose that XX is a geodesic metric space such that the Assouad-Nagata dimension dimAN(X)\dim_{\text{AN}}(X) of XX is finite. Suppose that one of the following is true:

  • ZZ is Lipschitz nn-connected.

  • XX is Lipschitz nn-connected, and if Xp,pPX_{p},p\in P are the connected components of XZX\smallsetminus Z, then the sets Hp=XpH_{p}=\partial X_{p} are Lipschitz nn-connected with uniformly bounded implicit constant.

Then ZZ is undistorted up to dimension n+1n+1.

In the applications in this paper, XX will be a CAT(0) space (either a symmetric space or a building), and ZZ will either be a horosphere of XX or the complement of a set of disjoint horoballs.

When XX is CAT(0), a theorem of Gromov [Gro83, Wen08] implies that the kkth-order Dehn function of XX grows at most as fast as V(k+1)/kV^{(k+1)/k} (i.e., if α\alpha is a Lipschitz kk-cycle in XX, there is a Lipschitz (k+1)(k+1)-chain β\beta in XX such that β=α\partial\beta=\alpha and

massβ(massα)(k+1)/k+c.\operatorname{mass}\beta\lesssim(\operatorname{mass}\alpha)^{(k+1)/k}+c.

Therefore,

Corollary 1.4.

If XX is CAT(0) and the hypotheses above hold, the kkth-order Dehn function of ZZ grows at most as fast as V(k+1)/kV^{(k+1)/k} for knk\leq n.

This bound is often sharp; for instance, if there is a rank-(k+1)(k+1) flat of XX contained in ZZ, then the kkth-order Dehn function of ZZ grows at least as fast as V(k+1)/kV^{(k+1)/k}.

We will apply Theorem 1.3 to find fillings in a family of solvable groups and in the Hilbert modular groups:

Theorem 1.5.

The group Sol2n1=n1n\operatorname{Sol}_{2n-1}=\mathbb{R}^{n-1}\ltimes\mathbb{R}^{n} is Lipschitz n1n-1-connected, is undistorted in (2)n(\mathbb{H}_{2})^{n} up to dimension nn, and its kkth-order Dehn function is asymptotic to V(k+1)/kV^{(k+1)/k} for k<nk<n.

This is a higher-dimensional version of a theorem of Gromov [Gro93, 5.A9] which states that Sol2n1\operatorname{Sol}_{2n-1} has quadratic Dehn function when n>1n>1. These bounds are sharp; there are nn-spheres in Sol2n1\operatorname{Sol}_{2n-1} with volume VV but filling volume exponential in VV, so the nnth order Dehn function of Sol2n1\operatorname{Sol}_{2n-1} is exponential [Gro93]. The same bounds apply to Hilbert modular groups:

Theorem 1.6.

If ΓSL2()n\Gamma\subset\operatorname{SL}_{2}(\mathbb{R})^{n} is a Hilbert modular group, then the kkth-order Dehn function of Γ\Gamma is asymptotic to V(k+1)/kV^{(k+1)/k} for k<nk<n.

We will also apply the methods of Theorem 1.3 to horospheres in euclidean buildings and to the SS-arithmetic groups considered in [BW11].

Let XX be a thick euclidean building and EXE\subset X be an apartment. Then the vertices of EE form a lattice, and if r:[0,)Er:[0,\infty)\to E is a geodesic ray, we say that rr has rational slope if it is parallel to a line segment connecting two vertices of EE. This condition is independent of the choice of EE, so if r:[0,)Xr:[0,\infty)\to X is a geodesic ray, we say it has rational slope if it has rational slope considered as a ray in some apartment EE. The boundary at infinity of XX consists of equivalence classes of geodesic rays, so if τX\tau\in X_{\infty} is a point in the boundary at infinity of XX, we say it has rational slope if one of the rays asymptotic to τ\tau has rational slope. In particular, if the isometry group of XX acts cocompactly on a horosphere centered at τ\tau, then τ\tau has rational slope.

Theorem 1.7.

Let XX be a thick euclidean building and let τX\tau\in X_{\infty} be a point in its boundary at infinity which has rational slope and is not contained in a factor of rank less than nn (in particular, XX has rank at least nn). Let ZZ be a horosphere in XX centered at τ\tau. Then ZZ is Lipschitz (n2)(n-2)-connected, undistorted in XX up to dimension n1n-1, and its kkth-order Dehn function grows at most as fast as V(k+1)/kV^{(k+1)/k} for kn2k\leq n-2.

ZZ is not (n1)(n-1)-connected, so the bound on kk is sharp. Indeed, for every r>0r>0, there is a map α:Sn1Z\alpha:S^{n-1}\to Z such that α\alpha is not null-homotopic in the rr-neighborhood of ZZ (see Lemma 4.15).

Note that if τ\tau does not have rational slope, then ZZ may be (n2)(n-2)-connected and locally Lipschitz (n2)(n-2)-connected but not Lipschitz (n2)(n-2)-connected. Cells of XX may intersect ZZ in arbitrarily small sets, and this can lead to arbitrarily small spheres which have small fillings in XX but filling volume 1\sim 1 in ZZ.

Theorem 1.7 is similar to Theorem 7.7 of [BW11], and gives a higher-order version of Theorem 1.1 of [Dru04] for buildings and products of buildings. (Though note that Theorem 1.1 of [Dru04] applies to \mathbb{R}-buildings as well as discrete buildings.)

The same methods lead to bounds on the higher-order Dehn functions of SS-arithmetic groups of KK-rank 1.

Theorem 1.8.

Let KK be a global function field, 𝐆\mathbf{G} be a noncommutative, absolutely almost simple KK-group of KK-rank 1, let SS be a finite set of pairwise inequivalent valuations on KK, and let XX be the associated euclidean building. Then the kkth-order Dehn function of the SS-arithmetic group G(𝒪S)G(\mathcal{O}_{S}) grows at most as fast as V(k+1)/kV^{(k+1)/k} for kdimX2k\leq\dim X-2.

This improves results of Bux and Wortman, who showed that G(𝒪S)G(\mathcal{O}_{S}) is of type FdimX1F_{\dim X-1} but not of type FdimXF_{\dim X} [BW11, BW07]. Bux and Wortman showed that horospheres in XX are (dimX2)(\dim X-2)-connected; Theorem 1.8 gives a quantitative proof of this fact.

Some possible other applications of Theorem 1.3 include the study of higher-order fillings in, for instance, metabelian groups, as in [dCT10], lattices of \mathbb{Q}-rank 1 in semisimple groups, as in [Dru04], and SS-arithmetic lattices when |S||S| is large, as in [BEW].

Notational conventions: If ff and gg are expressions, we will write fgf\lesssim g if there is some constant cc such that fcgf\leq cg. We write fgf\sim g if there is some constant cc such that c1fcgc^{-1}\leq f\leq cg. When we wish to emphasize that cc depends on xx and yy, we write fx,ygf\lesssim_{x,y}g or fx,ygf\sim_{x,y}g. We give SkS^{k} the round metric, scaled so that diamSk=1\operatorname{diam}S^{k}=1, and we define the standard kk-simplex to be the equilateral euclidean kk-simplex, scaled to have diameter 1.

Acknowledgements: This work was supported by a Discovery Grant from the Natural Sciences and Engineering Research Council of Canada and by the Connaught Fund, University of Toronto. The author would like to thank MSRI and the organizers of the 2011 Quantitative Geometry program for their hospitality, and would like to thank Cornelia Druţu, Enrico Leuzinger, Romain Tessera, and Kevin Wortman for helpful discussions and suggestions.

2. Building fillings from simplices

The proof of Theorem 1.3 is based on the proof of a theorem of Lang and Schlichenmaier. Lang and Schlichenmaier proved:

Theorem 2.1.

Suppose that ZXZ\subset X is a nonempty closed set and that dimANXm<\dim_{\text{AN}}{X}\leq m<\infty. If YY is Lipschitz (m1)(m-1)-connected, then there is a CC such that any Lipschitz map f:ZYf:Z\to Y extends to a map f¯:XY\bar{f}:X\to Y with Lip(f¯)CLip(f)\operatorname{Lip}(\bar{f})\leq C\operatorname{Lip}(f).

Here, dimAN(X)\dim_{\text{AN}}(X) is the Assouad-Nagata dimension of XX. The Assouad-Nagata dimension of XX is the smallest integer such that there is a cc such that for all s>0s>0, there is a covering s\mathcal{B}_{s} of XX by sets of diameter at most cscs (a cscs-bounded covering) such that any set with diameter s\leq s intersects at most n+1n+1 sets in the cover (i.e., s\mathcal{B}_{s} has ss-multiplicity at most n+1n+1).

One consequence of Theorem 2.1 is that if ZZ is Lipschitz nn-connected for n=dimAN(X)n=\dim_{\text{AN}}(X), then the identity map ZZZ\to Z can be extended to a Lipschitz map f¯:XZ\bar{f}:X\to Z and ZZ is a Lipschitz retract of XX. Consequently, if α\alpha is a (k1)(k-1)-cycle in ZZ and β\beta is a chain in XX with boundary α\alpha, then f¯(β)\bar{f}_{\sharp}(\beta) is a chain in ZZ with boundary α\alpha, so

FVZk(α)CkFVXk(α),\operatorname{FV}^{k}_{Z}(\alpha)\leq C^{k}\operatorname{FV}_{X}^{k}(\alpha),

and ZZ is undistorted in XX up to dimension nn. Theorem 1.3 claims that the same is true under the weaker condition that XX has finite Assouad-Nagata dimension.

Before we sketch the proof of Theorem 1.3, we need the notion of a quasi-conformal complex. We define a riemannian simplicial complex to be a simplicial complex with a metric which gives each simplex the structure of a riemannian manifold with corners. We say that such a complex is quasi-conformal (or that the complex is a QC complex) if there is a cc such that the riemannian metric on each simplex is cc-bilipschitz equivalent to a scaling of the standard simplex.

QC complexes are a compromise between the rigidity of simplicial complexes and the freedom of riemannian simplicial complexes. A key feature of simplicial complexes is that curves and cycles can be approximated by simplicial curves and cycles. This is not true in riemannian simplicial complexes, but it holds in QC complexes.

Specifically, a version of the Federer-Fleming deformation theorem holds in QC complexes. Recall that the Federer-Fleming theorem for simplicial complexes states that any Lipschitz cycle aa in a simplicial complex can be approximated by a simplicial cycle P(a)P(a) whose mass is comparable to the mass of aa. We will use the following variation of the Federer-Fleming theorem:

Theorem 2.2.

Let Σ\Sigma be a finite-dimensional scaled simplicial complex, that is, a simplicial complex where each simplex is given the metric of the standard simplex of diameter ss. There is a constant cc depending on dimΣ\dim\Sigma such that if aCkLip(Σ)a\in C^{\text{Lip}}_{k}(\Sigma) is a Lipschitz kk-cycle, then there are P(a)Ckcell(X)P(a)\in C^{\text{cell}}_{k}(X) and Q(a)Ck+1Lip(X)Q(a)\in C^{\text{Lip}}_{k+1}(X) such that

  1. (1)

    a=P(a)\partial a=\partial P(a)

  2. (2)

    Q(a)=aP(a)\partial Q(a)=a-P(a)

  3. (3)

    massP(a)cmass(a)\operatorname{mass}P(a)\leq c\cdot\operatorname{mass}(a)

  4. (4)

    massQ(a)csmass(a)\operatorname{mass}Q(a)\leq cs\cdot\operatorname{mass}(a)

A proof of this theorem when s=1s=1 can be found in [ECH+92]. A simple scaling argument proves the general case. Note that, while the bound on massQ(a)\operatorname{mass}Q(a) depends on the size of the simplices, the bound on massP(a)\operatorname{mass}P(a) does not.

Because the bound on massP(a)\operatorname{mass}P(a) is independent of the size of the simplices in the complex, the following version of Theorem 2.2 holds for a QC complex:

Theorem 2.3.

Let Σ\Sigma be a QC complex. There is a constant cc depending on dimΣ\dim\Sigma such that if aCkLip(Σ)a\in C^{\text{Lip}}_{k}(\Sigma) is a Lipschitz kk-cycle, then there is a P(a)Ckcell(X)P(a)\in C^{\text{cell}}_{k}(X) such that a=P(a)\partial a=\partial P(a) and massP(a)cmass(a)\operatorname{mass}P(a)\leq c\cdot\operatorname{mass}(a).

Now we will sketch a proof of Theorem 1.3. Note that this sketch is incorrect due to some technical issues; we will fix these issues in the actual proof. In the proof of Theorem 1.5 of [LS05], Lang and Schlichenmaier show that, if dimAN(X)<\dim_{\text{AN}}(X)<\infty, there are a>0a>0, 0<b<10<b<1 and a cover =(Bi)iI0\mathcal{B}=(B_{i})_{i\in I_{0}} of XZX\setminus Z by subsets of XZX\setminus Z such that:

  1. (1)

    diamBiad(Bi,Z)\operatorname{diam}B_{i}\leq ad(B_{i},Z) for every iI0i\in I_{0}

  2. (2)

    every set DXZD\subset X\setminus Z with diamDbd(D,Z)\operatorname{diam}D\leq bd(D,Z) meets at most dimAN(X)+1\dim_{\text{AN}}(X)+1 members of (Bi)iI0(B_{i})_{i\in I_{0}}.

They then define functions σi:XZ\sigma_{i}:X\setminus Z\to\mathbb{R},

σi(x)=max{0,δd(Bi,Z)d(x,Bi)},\sigma_{i}(x)=\max\{0,\delta d(B_{i},Z)-d(x,B_{i})\},

where δ=b/(2(b+1))\delta=b/(2(b+1)), and show that these have the property that for any xx, there are no more than dimAN(X)+1\dim_{\text{AN}}(X)+1 values of ii for which σi(x)>0\sigma_{i}(x)>0. Using these σi\sigma_{i}, they construct a Lipschitz map g:XZΣ0g:X\setminus Z\to\Sigma_{0}, where Σ0\Sigma_{0} is the nerve of the supports of the σi\sigma_{i}. One can give Σ0\Sigma_{0} the structure of a QC complex so that if Δ\Delta is a simplex of Σ0\Sigma_{0} with a vertex corresponding to σi\sigma_{i}, then diamΔdiamsuppσi\operatorname{diam}\Delta\sim\operatorname{diam}\operatorname{supp}\sigma_{i}. Since the diameter of suppσi\operatorname{supp}\sigma_{i} is proportional to d(σi,Z)d(\sigma_{i},Z), this means that the parts of Σ0\Sigma_{0} which are close to ZZ are given a fine triangulation and the parts of Σ0\Sigma_{0} which are far from ZZ are given a coarse triangulation.

Since ZZ is Lipschitz nn-connected, one can construct a Lipschitz map h:Σ0(n+1)Zh:\Sigma_{0}^{(n+1)}\to Z, where Σ0(n+1)\Sigma_{0}^{(n+1)} is the (n+1)(n+1)-skeleton of Σ0\Sigma_{0}. Then, if α\alpha is an nn-cycle in ZZ, it has a filling β\beta in XX. We can use the Federer-Fleming theorem to approximate g(β)g_{\sharp}(\beta) by some simplicial (n+1)(n+1)-chain P(β)P(\beta) which lies in Σ0(n+1)\Sigma_{0}^{(n+1)}. The pushforward of P(β)P(\beta) under hh will then be a filling of α\alpha.

The problem with this argument is twofold. First, since gg is only defined on XZX\setminus Z, we can’t define g(β)g_{\sharp}(\beta) without extending gg to ZZ. We could define an appropriate metric on the disjoint union Σ0Z\Sigma_{0}\amalg Z and a map XΣ0ZX\to\Sigma_{0}\amalg Z, but this is no longer a simplicial complex. Second, since the cells of Σ\Sigma get arbitrarily small close to ZZ, P(β)P(\beta) may be an infinite sum of cells of Σ\Sigma.

We know of two ways to fix this issue. First, one can make sense of infinite sums of cells of Σ\Sigma by introducing Lipschitz currents [AK00]. The set of Lipschitz currents is a completion of the set of Lipschitz chains, and the P(β)P(\beta) defined above is a current in Σ0Z\Sigma_{0}\amalg Z. Its pushforward is then a filling of α\alpha. Second, we can change the construction of Σ0\Sigma_{0} to avoid the problem. We take this approach in the rest of this section.

All the constants and all the implicit constants in \lesssim and \sim in this section will depend on X,ZX,Z, and nn.

First, we construct a QC complex Σ\Sigma which approximates XX. This complex will have geometry similar to Σ0\Sigma_{0} on XZX\setminus Z and it will have ϵ\epsilon-small simplices on ZZ. For t>0t>0, let Nt(Z)XN_{t}(Z)\subset X be the tt-neighborhood of ZZ.

Lemma 2.4.

There are a,b,γ>0a,b,\gamma>0 such that if ϵ>0\epsilon>0 and δ=b/(2(b+1))\delta=b/(2(b+1)), there is a covering 𝒟\mathcal{D} of XX by sets DkD_{k}, kKk\in K and functions r:Kr:K\to\mathbb{R}, τk:X\tau_{k}:X\to\mathbb{R}

r(k)=max{δd(Dk,Z),ϵ}r(k)=\max\{\delta d(D_{k},Z),\epsilon\}
τk(x)=max{0,r(k)d(x,Dk)}\tau_{k}(x)=\max\{0,r(k)-d(x,D_{k})\}

such that for any kKk\in K,

  1. (1)

    diamDkr(k),\operatorname{diam}D_{k}\lesssim r(k),

  2. (2)

    d(Dk,Z)r(k),d(D_{k},Z)\lesssim r(k),

  3. (3)

    if ρ=ϵδ(1+a)\rho=\epsilon\delta(1+a) and d(Dk,Z)ρd(D_{k},Z)\geq\rho, then suppτk\operatorname{supp}\tau_{k} is contained in a connected component of XZX\smallsetminus Z,

  4. (4)

    the cover of XX by the sets suppτk\operatorname{supp}\tau_{k} has multiplicity at most 2dimAN(X)+22\dim_{\text{AN}}(X)+2, and

  5. (5)

    if suppτksuppτk\operatorname{supp}\tau_{k}\cap\operatorname{supp}\tau_{k^{\prime}}\neq\emptyset, then

    γ1r(k)r(k)γr(k).\gamma^{-1}r(k^{\prime})\leq r(k)\leq\gamma r(k^{\prime}).
Proof.

Let a>0a>0, 0<b<10<b<1, and =(Bi)iI0\mathcal{B}=(B_{i})_{i\in I_{0}} be as in the Lang-Schlichenmaier construction above. Let We may assume that each BiB_{i} is contained in a connected component of XZX\smallsetminus Z. Let ρ=ϵδ(1+a)\rho=\epsilon\delta(1+a), and let II0I\subset I_{0} be the set

I:={iI0BiNρ(Z)}.I:=\{i\in I_{0}\mid B_{i}\not\subset N_{\rho}(Z)\}.

Then iIBiXNρ(Z)\bigcup_{i\in I}B_{i}\supset X\setminus N_{\rho}(Z). Since dimAN(X)\dim_{\text{AN}}(X)\leq\infty, we can let 𝒞={Cj}jJ\mathcal{C}=\{C_{j}\}_{j\in J} be a 2c0ϵ2c_{0}\epsilon-bounded covering of Nρ(Z)N_{\rho}(Z) with 2ϵ2\epsilon-multiplicity at most dimAN(X)+1\dim_{\text{AN}}(X)+1, where c0c_{0} is the constant in the definition of dimAN(X)\dim_{\text{AN}}(X). Let 𝒟=𝒞{Bi}iI\mathcal{D}=\mathcal{C}\cup\{B_{i}\}_{i\in I} and let K=IJK=I\amalg J.

Conditions (1) and (2) are easy to check. For (3), note that if d(Dk,Z)ρd(D_{k},Z)\geq\rho, then kIk\in I, so Di=BiD_{i}=B_{i} lies in a single connected component of XZX\setminus Z, and suppτi\operatorname{supp}\tau_{i} lies in the same component. For (4), note that if iIi\in I, then τi=σi\tau_{i}=\sigma_{i}, so the cover {suppτi}iI\{\operatorname{supp}\tau_{i}\}_{i\in I} has multiplicity at most dimAN(X)+1\dim_{\text{AN}}(X)+1. Likewise, if xsuppτjx\in\operatorname{supp}\tau_{j} for some jJj\in J, then CjB(x,ϵ)C_{j}\cap B(x,\epsilon)\neq\emptyset, where B(x,ϵ)B(x,\epsilon) is the closed ball of radius ϵ\epsilon around xx. Since 𝒞\mathcal{C} has bounded 2ϵ2\epsilon-multiplicity, this can be true for only dimAN(X)+1\dim_{\text{AN}}(X)+1 values of jj.

To check (5), suppose that suppτksuppτk\operatorname{supp}\tau_{k}\cap\operatorname{supp}\tau_{k^{\prime}}\neq\emptyset. If r(k)=ϵr(k^{\prime})=\epsilon, then r(k)r(k)r(k^{\prime})\leq r(k). Otherwise, r(k)=δd(Dk,Z)r(k^{\prime})=\delta d(D_{k^{\prime}},Z). But d(Dk,Dk)r(k)d(D_{k},D_{k^{\prime}})\lesssim r(k) and diamDkr(k)\operatorname{diam}D_{k}\lesssim r(k), so d(Dk,Z)r(k)d(D_{k^{\prime}},Z)\lesssim r(k), and r(k)r(k)r(k^{\prime})\lesssim r(k). By symmetry, r(k)r(k)r(k)\sim r(k^{\prime}). ∎

Let Σ\Sigma be the nerve of the cover {suppτk}kK\{\operatorname{supp}\tau_{k}\}_{k\in K}, with vertex set {vk}kK\{v_{k}\}_{k\in K} and let s:Σs:\Sigma\to\mathbb{R} is the function such that s(vk)=r(k)s(v_{k})=r(k) and ss is linear on each simplex of Σ\Sigma. Define a riemannian metric xcx_{c} on each simplex of Σ\Sigma by letting dxc2=s2dx2dx_{c}^{2}=s^{2}\;dx^{2}. If S=vk1,vknS=\langle v_{k_{1}},\dots v_{k_{n}}\rangle is a simplex of Σ\Sigma, then ss varies between γ1r(k1)\gamma^{-1}r(k_{1}) and γr(k1)\gamma r(k_{1}) on SS, so this metric makes Σ\Sigma a QC complex.

Lemma 2.5.

There is a Lipschitz map g:XΣg:X\to\Sigma with Lipschitz constant c1c_{1} independent of ϵ\epsilon. Furthermore, if xsuppτkx\in\operatorname{supp}\tau_{k} for some kKk\in K, then g(x)g(x) is in the star of vkv_{k}.

Proof.

Consider the infinite simplex

ΔK:={p:K[0,1]p1=1}\Delta_{K}:=\{p:K\to[0,1]\mid\|p\|_{1}=1\}

with vertex set KK, so that Σ\Sigma is a subcomplex of ΔK\Delta_{K}. Let

g(x)(k)=τk(x)τ¯(x),g(x)(k)=\frac{\tau_{k}(x)}{\bar{\tau}(x)},

where τ¯(x)=kKτk(x)\bar{\tau}(x)=\sum_{k\in K}\tau_{k}(x). The image of gg then lies in Σ\Sigma, and we can consider gg as a function XΣX\to\Sigma.

It remains to show that gg is Lipschitz with respect to the QC metric on Σ\Sigma. Since XX is geodesic, it suffices to show that if x,yXx,y\in X and d(x,y)<δ2ϵ<ϵ,d(x,y)<\delta^{2}\epsilon<\epsilon, then d(g(x),g(y))d(x,y)d(g(x),g(y))\lesssim d(x,y). Let SS and TT be the minimal simplices of Σ\Sigma which contain g(x)g(x) and g(y)g(y) respectively. First, we claim that SS and TT share some vertex vmv_{m}.

Let ρ=ϵδ(1+a)\rho=\epsilon\delta(1+a) as above. If d(x,Z)<ρd(x,Z)<\rho, then there is some jJj\in J such that xCjx\in C_{j} and τj(x)=ϵ\tau_{j}(x)=\epsilon. Since τj\tau_{j} is 1-Lipschitz, τj(y)>0\tau_{j}(y)>0, so we can let m=jm=j. Otherwise, if d(x,Z)ρd(x,Z)\geq\rho, then there is some iIi\in I such that xBix\in B_{i}. We have

d(x,Z)diam(Bi)+d(Bi,Z)(a+1)d(Bi,Z),d(x,Z)\leq\operatorname{diam}(B_{i})+d(B_{i},Z)\leq(a+1)d(B_{i},Z),

so τi(x)=δd(Bi,Z)δ2ϵ\tau_{i}(x)=\delta d(B_{i},Z)\geq\delta^{2}\epsilon, and τi(y)>0\tau_{i}(y)>0 as desired. We let m=im=i.

Since SS and TT share vmv_{m}, the value of ss on STS\cup T is at most γr(m)\gamma r(m), and

d(g(x),g(y))\displaystyle d(g(x),g(y)) γr(m)k(ST)(0)|τk(x)τ¯(x)τk(y)τ¯(y)|\displaystyle\leq\gamma r(m)\sum_{k\in(S\cup T)^{(0)}}\left|\frac{\tau_{k}(x)}{\bar{\tau}(x)}-\frac{\tau_{k}(y)}{\bar{\tau}(y)}\right|
γr(m)k(ST)(0)|τk(x)τ¯(x)τk(y)τ¯(x)|+|τk(y)τ¯(x)τk(y)τ¯(y)|\displaystyle\leq\gamma r(m)\sum_{k\in(S\cup T)^{(0)}}\left|\frac{\tau_{k}(x)}{\bar{\tau}(x)}-\frac{\tau_{k}(y)}{\bar{\tau}(x)}\right|+\left|\frac{\tau_{k}(y)}{\bar{\tau}(x)}-\frac{\tau_{k}(y)}{\bar{\tau}(y)}\right|
γr(m)k(ST)(0)1τ¯(x)(|τk(x)τk(y)|+τk(y)τ¯(y)|τ¯(x)τ¯(y)|)\displaystyle\leq\gamma r(m)\sum_{k\in(S\cup T)^{(0)}}\frac{1}{\bar{\tau}(x)}\left(|\tau_{k}(x)-\tau_{k}(y)|+\frac{\tau_{k}(y)}{\bar{\tau}(y)}|\bar{\tau}(x)-\bar{\tau}(y)|\right)
γ(2dimΣ+1)(2dimΣ+2)r(m)τ¯(x)d(x,y)\displaystyle\leq\gamma(2\dim\Sigma+1)(2\dim\Sigma+2)\frac{r(m)}{\bar{\tau}(x)}d(x,y)

Furthermore, if xDmx\in D_{m^{\prime}}, then

τ¯(x)r(m)γ1r(m),\bar{\tau}(x)\geq r(m^{\prime})\geq\gamma^{-1}r(m),

so gg has Lipschitz constant at most

c1=γ2(2dimΣ+1)(2dimΣ+2).c_{1}=\gamma^{2}(2\dim\Sigma+1)(2\dim\Sigma+2).

Next, we construct a map h:Σ(n+1)Zh:\Sigma^{(n+1)}\to Z on the (n+1)(n+1)-skeleton of Σ\Sigma. If Δ\Delta is a simplex of Σ\Sigma, denote its vertex set by 𝒱(Δ)\mathcal{V}(\Delta).

Lemma 2.6.

For any ϵ>0\epsilon^{\prime}>0, there is a Lipschitz map h(0):Σ(0)Zh^{(0)}:\Sigma^{(0)}\to Z with Lipschitz constant independent of ϵ\epsilon which satisfies:

  1. (1)

    d(h(0)(vj),Cj)ϵd(h^{(0)}(v_{j}),C_{j})\lesssim\epsilon for every jJj\in J,

  2. (2)

    if Xp,pPX_{p},p\in P are the connected components of XZX\smallsetminus Z and

    Hp(ϵ)={xXd(x,Xp)ϵ}Z,H_{p}(\epsilon^{\prime})=\{x\in X\mid d(x,X_{p})\leq\epsilon^{\prime}\}\cap Z,

    then for any simplex ΔΣ\Delta\subset\Sigma, we either have diamh(0)(𝒱(Δ))ϵ\operatorname{diam}h^{(0)}(\mathcal{V}(\Delta))\lesssim\epsilon (if Δ\Delta has a vertex of the form vjv_{j} for some jJj\in J) or h(0)(𝒱(Δ))Hp(ϵ)h^{(0)}(\mathcal{V}(\Delta))\subset H_{p}(\epsilon^{\prime}) for some pPp\in P (otherwise).

Proof.

For each vertex vkΣv_{k}\in\Sigma, choose a point zkZz_{k}\in Z such that d(zk,Dk)<d(Z,Dk)+ϵH/2d(z_{k},D_{k})<d(Z,D_{k})+\epsilon_{H}/2, and let h(0)(vk)=zkh^{(0)}(v_{k})=z_{k}. If kJk\in J, then d(Z,Dk)ϵd(Z,D_{k})\lesssim\epsilon, so d(zk,Dk)ϵd(z_{k},D_{k})\lesssim\epsilon and property (1) holds. We claim that this map is Lipschitz. Suppose that v,wv,w are vertices of Σ\Sigma. Then there is a path γ:[0,1]Σ\gamma:[0,1]\to\Sigma between them of length (γ)2d(v,w)\ell(\gamma)\leq 2d(v,w), and the Federer-Fleming theorem implies that this can be approximated by a path γ:[0,1]Σ(1)\gamma^{\prime}:[0,1]\to\Sigma^{(1)} in the 1-skeleton of Σ\Sigma with (γ)(γ)\ell(\gamma^{\prime})\lesssim\ell(\gamma). So, to check that h(0)h^{(0)} is Lipschitz, it suffices to show that if vkv_{k} and vkv_{k^{\prime}} are connected by an edge ee, then d(zk,zk)(e)d(z_{k},z_{k^{\prime}})\lesssim\ell(e).

We may assume that r(k)r(k)r(k)\geq r(k^{\prime}), so (e)γ1r(k)\ell(e)\geq\gamma^{-1}r(k). Then we can bound d(zk,zk)d(z_{k},z_{k^{\prime}}) by

d(zk,zk)d(zk,Dk)+diam(Dk)+d(Dk,Dk)+diam(Dk)+d(Dk,zk)d(z_{k},z_{k^{\prime}})\leq d(z_{k},D_{k})+\operatorname{diam}(D_{k})+d(D_{k},D_{k^{\prime}})+\operatorname{diam}(D_{k^{\prime}})+d(D_{k^{\prime}},z_{k^{\prime}})

Each term on the right is r(k)\lesssim r(k). For each term except d(Dk,Dk)d(D_{k},D_{k^{\prime}}), this follows from the remarks after the definition of SS. To bound d(Dk,Dk)d(D_{k},D_{k^{\prime}}), note that since there is an edge from vkv_{k} to vkv_{k^{\prime}}, there is a wsuppτksuppτkw\in\operatorname{supp}\tau_{k}\cap\operatorname{supp}\tau_{k^{\prime}}. Then d(w,Dk)<r(k)d(w,D_{k})<r(k) and d(w,Dk)<r(k)d(w,D_{k^{\prime}})<r(k), so d(Dk,Dk)2r(k)d(D_{k},D_{k^{\prime}})\leq 2r(k). Therefore, h(0)h^{(0)} is Lipschitz.

It remains to check property (2). Let Δ=vk0,,vkn\Delta=\langle v_{k_{0}},\dots,v_{k_{n}}\rangle be a simplex of Σ\Sigma and suppose that kiJk_{i}\in J for some ii. Then r(ki)ϵr(k_{i})\lesssim\epsilon, so diamΔϵ\operatorname{diam}\Delta\lesssim\epsilon, and therefore, diamh(0)(𝒱(Δ))ϵ\operatorname{diam}h^{(0)}(\mathcal{V}(\Delta))\lesssim\epsilon.

Otherwise, kiIk_{i}\in I for all ii. Then there is some pPp\in P such that suppτkiXp\operatorname{supp}\tau_{k_{i}}\subset X_{p} for all ii, and h(0)(𝒱(Δ))Hp(ϵ)h^{(0)}(\mathcal{V}(\Delta))\subset H_{p}(\epsilon^{\prime}). ∎

If ϵ>0\epsilon>0 and nn are such that whenever knk\leq n and τ:SkZ\tau:S^{k}\to Z is a map with Lipτϵ\operatorname{Lip}\tau\leq\epsilon, there is an extension τ¯:Dk+1Z\bar{\tau}:D^{k+1}\to Z with Lipτ¯Lipτ\operatorname{Lip}\bar{\tau}\lesssim\operatorname{Lip}\tau, we say that YY is ϵ\epsilon-locally Lipschitz nn-connected.

Lemma 2.7.

If XX and ZZ satisfy the conditions of Theorem 1.3 and ϵ\epsilon is sufficiently small, then there is a Lipschitz extension h:Σ(n+1)Zh:\Sigma^{(n+1)}\to Z with Lipschitz constant independent of ϵ\epsilon such that d(h(g(z)),z)ϵd(h(g(z)),z)\lesssim\epsilon for every zZz\in Z.

Proof.

In this proof, it will be convenient to let SkS^{k} be the boundary of the standard (k+1)(k+1)-simplex and DkD^{k} be the standard kk-simplex. If t>0t>0, we let tSktS^{k} and tDktD^{k} be scalings of SkS^{k} and DkD^{k}. If a space YY is Lipschitz nn-connected, there is a cc such that if knk\leq n, any Lipschitz map τ:SkY\tau:S^{k}\to Y can be extended to a Lipschitz map τ¯:Dk+1Y\bar{\tau}:D^{k+1}\to Y with Lipτ¯cLipτ\operatorname{Lip}\bar{\tau}\leq c\operatorname{Lip}\tau. By scaling, any Lipschitz map τ:tSkY\tau:tS^{k}\to Y can be extended to a Lipschitz map τ¯:tDk+1Y\bar{\tau}:tD^{k+1}\to Y with Lipτ¯cLipτ\operatorname{Lip}\bar{\tau}\leq c\operatorname{Lip}\tau

If ZZ is Lipschitz nn-connected, then we can use Lipschitz nn-connectivity to extend h(0)h^{(0)}. That is, if we have already defined hh on Σ(k)\Sigma^{(k)} and ΔΣ\Delta\subset\Sigma is a (k+1)(k+1)-simplex, then the Riemannian metric on Δ\Delta is bilipschitz equivalent to s(x)Dk+1s(x)D^{k+1} for any xΔx\in\Delta. Since h|Δh|_{\partial\Delta} is a Lipschitz map of a kk-sphere, we can extend hh over Δ\Delta, and the extension satisfies LiphLiph(0)\operatorname{Lip}h\lesssim\operatorname{Lip}h^{(0)}.

If ZZ is not Lipschitz nn-connected, we need a more careful approach. By hypothesis, XX is Lipschitz nn-connected; let c>0c>0 be the constant in the definition of Lipschitz nn-connectivity.

Let ϵ=ϵH/c\epsilon^{\prime}=\epsilon_{H}/c and let knk\leq n. If τ:SkZ\tau:S^{k}\to Z is a map with Lipτϵ\operatorname{Lip}\tau\leq\epsilon^{\prime}, we claim that τ\tau can be extended to a Lipschitz map on Dk+1D^{k+1}. If τ(Sk)Hp(ϵH)\tau(S^{k})\subset H_{p}(\epsilon_{H}) for some pPp\in P, then we can extend τ\tau to Dk+1D^{k+1} using the Lipschitz nn-connectivity of Hp(ϵH)H_{p}(\epsilon_{H}). Otherwise, there is some xSkx\in S^{k} such that d(τ(x),XZ)>ϵHd(\tau(x),X\setminus Z)>\epsilon_{H}. Since diamτ¯0(Dk+1)ϵH\operatorname{diam}\bar{\tau}_{0}(D^{k+1})\leq\epsilon_{H}, the image of τ¯0\bar{\tau}_{0} is contained in ZZ. Therefore, ZZ is ϵ\epsilon^{\prime}-locally Lipschitz nn-connected.

If ΔΣ\Delta\subset\Sigma is a simplex, we say that it is coarse if all its vertices are of the form viv_{i} for iIi\in I. We say that it is fine if it has a vertex of the form vjv_{j} for some jJj\in J; all fine simplices have diameter ϵ\lesssim\epsilon and all coarse ones have diameter ϵ\gtrsim\epsilon. By the previous lemma, we can choose h(0)h^{(0)} so that for every coarse simplex Δ\Delta, there is some pPp\in P such that h(0)(𝒱(Δ))Hp(ϵH)h^{(0)}(\mathcal{V}(\Delta))\subset H_{p}(\epsilon_{H}). If ΣcΣ\Sigma_{c}\subset\Sigma is the subcomplex consisting of coarse simplices, we can extend h(0)h^{(0)} to a map hc:Σ(0Σc(n+1)Zh_{c}:\Sigma^{(0}\cup\Sigma^{(n+1)}_{c}\to Z by induction; if hc|Δh_{c}|_{\partial\Delta} is defined, then hc(Δ)Hp(ϵH)h_{c}(\partial\Delta)\subset H_{p}(\epsilon_{H}) for some pPp\in P. We extend hch_{c} over Δ\Delta using the Lipschitz nn-connectivity of Hp(ϵH)H_{p}(\epsilon_{H}). The Lipschitz constant of hch_{c} is bounded independently of ϵ\epsilon.

Again by the previous lemma, we may choose ϵ\epsilon sufficiently small that any fine simplex has diameter ϵ/Liphc\ll\epsilon^{\prime}/\operatorname{Lip}h_{c}. We can then extend hch_{c} over the fine simplices of Σ\Sigma using the local Lipschitz connectivity of ZZ to get the desired map hh.

In either case, if zZz\in Z, then zsuppτkz\in\operatorname{supp}\tau_{k} only if kJk\in J. In particular, g(z)g(z) is contained in a fine simplex of diameter ϵ\lesssim\epsilon and d(z,zi)ϵd(z,z_{i})\lesssim\epsilon, so

d(h(g(z)),z)d(h(g(z)),h(vi))+d(zi,z)ϵd(h(g(z)),z)\leq d(h(g(z)),h(v_{i}))+d(z_{i},z)\lesssim\epsilon

as desired. ∎

Therefore, hgh\circ g has small displacement. To complete the proof of Theorem 1.3, we will need a lemma concerning such maps:

Lemma 2.8.

Suppose that mnm\leq n, that α\alpha is a Lipschitz mm-cycle in ZZ, that ZZ is ϵ0\epsilon_{0}-locally Lipschitz nn-connected, and that C>0C>0. For any ϵ>0\epsilon>0, there is a δ>0\delta>0 such that if f:ZZf:Z\to Z is a CC-Lipschitz map with displacement δ\leq\delta (i.e., d(f(z),z)δd(f(z),z)\leq\delta for all zZz\in Z), then

FVZm+1(f(α)α)ϵ.\operatorname{FV}_{Z}^{m+1}(f_{\sharp}(\alpha)-\alpha)\leq\epsilon.
Proof.

Since ZZ is locally Lipschitz nn-connected, if MM is a simplicial (m+1)(m+1)-complex, NN is a subcomplex, and f:NZf:N\to Z is a map with sufficiently small Lipschitz constant, then there is an extension f¯:MZ\bar{f}:M\to Z with Lipschitz constant Lip(f)\sim\operatorname{Lip}(f). Write α\alpha as a sum α=i=1kαi\alpha=\sum_{i=1}^{k}\alpha_{i} of Lipschitz maps αi:ΔmZ\alpha_{i}:\Delta^{m}\to Z. Let LL be the maximum Lipschitz constant of the αi\alpha_{i}’s. In the following calculations, all our implicit constants will depend on kk, nn, ZZ, CC, and LL. We claim that

FVZm+1(f(α)α)δ.\operatorname{FV}_{Z}^{m+1}(f_{\sharp}(\alpha)-\alpha)\lesssim\delta.

First, we can subdivide Δm\Delta^{m} into δm\lesssim\delta^{-m} simplices each with diameter δ/L\leq\delta/L. We can use this subdivision to replace α\alpha with a sum α=i=1kαi\alpha^{\prime}=\sum_{i=1}^{k^{\prime}}\alpha^{\prime}_{i} where kδmk^{\prime}\lesssim\delta^{-m} and each αi:ΔmZ\alpha^{\prime}_{i}:\Delta^{m}\to Z has Lipschitz constant at most δ\delta.

There is a simplicial mm-complex AA with at most kk^{\prime} top-dimensional faces, a simplicial cycle ω\omega on AA, and a map g:AZg:A\to Z with Lip(g)δ\operatorname{Lip}(g)\leq\delta such that the restriction of gg to each top-dimensional face of AA is one of the αi\alpha^{\prime}_{i}’s and g(ω)=αg_{\sharp}(\omega)=\alpha^{\prime}. Define r0:A×{0,1}Zr_{0}:A\times\{0,1\}\to Z by letting r0|A×0=gr_{0}|_{A\times 0}=g and r0|A×1=fgr_{0}|_{A\times 1}=f\circ g. Then Lip(r0)δ\operatorname{Lip}(r_{0})\lesssim\delta, and if δ\delta is sufficiently small, we can extend it to a Lipschitz map r:A×[0,1]Zr:A\times[0,1]\to Z with LiprLipr0\operatorname{Lip}r\sim\operatorname{Lip}r_{0}. This is a homotopy from gg to fgf\circ g, so the push-forward of ω×[0,1]\omega\times[0,1] is a filling of f(α)αf_{\sharp}(\alpha)-\alpha with mass

massr(ω×[0,1])kδm+1δ\operatorname{mass}r_{\sharp}(\omega\times[0,1])\lesssim k^{\prime}\delta^{m+1}\lesssim\delta

as desired. ∎

Proof of Theorem 1.3.

Suppose that α\alpha is a (m1)(m-1)-cycle in ZZ and β\beta is a mm-chain filling it. Let ΣJ\Sigma_{J} be the subcomplex of Σ\Sigma spanned by the vertices vj,jJv_{j},j\in J. Then g(Z)ΣJg(Z)\subset\Sigma_{J}, and g(α)g_{\sharp}(\alpha) is a cycle in ΣJ\Sigma_{J} with mass Lip(g)m1massα\leq\operatorname{Lip}(g)^{m-1}\operatorname{mass}\alpha. Each simplex of ΣJ\Sigma_{J} has diameter ϵ\sim\epsilon, so by Thm. 2.2, there is a c3>0c_{3}>0 depending only on XX, a simplicial cycle Pα:=PΣJ(g(α))P_{\alpha}:=P_{\Sigma_{J}}(g_{\sharp}(\alpha)) approximating g(α)g_{\sharp}(\alpha), and a chain Qα:=QΣJ(g(α))Q_{\alpha}:=Q_{\Sigma_{J}}(g_{\sharp}(\alpha)) such that massQαc3ϵmass(α)\operatorname{mass}Q_{\alpha}\leq c_{3}\epsilon\operatorname{mass}(\alpha) and Qα=Pαg(α)\partial Q_{\alpha}=P_{\alpha}-g_{\sharp}(\alpha).

Then g(β)+Qαg_{\sharp}(\beta)+Q_{\alpha} is a mm-chain in Σ\Sigma with boundary PαP_{\alpha} and mass

mass(g(β)+Qα)Lip(g)mmassβ+c3ϵmass(α).\operatorname{mass}(g_{\sharp}(\beta)+Q_{\alpha})\leq\operatorname{Lip}(g)^{m}\operatorname{mass}\beta+c_{3}\epsilon\operatorname{mass}(\alpha).

Thm. 2.3 lets us approximate this by a chain

Pβ:=PΣ(g(β)+Qα)P_{\beta}:=P_{\Sigma}(g_{\sharp}(\beta)+Q_{\alpha})

with boundary PαP_{\alpha}.

By Lemma 2.8, if ϵ0>0\epsilon_{0}>0, then for ϵ\epsilon sufficiently small, there is a Lipschitz m+1m+1-chain RR in ZZ such that

R=α(hg)(α)\partial R=\alpha-(h\circ g)_{\sharp}(\alpha)

and massRϵ0\operatorname{mass}R\leq\epsilon_{0}. Let

B=Rh(Qα)h(Pβ)B=R-h_{\sharp}(Q_{\alpha})-h_{\sharp}(P_{\beta})

Then B=α\partial B=\alpha and

massBmassβ+ϵmass(α)+ϵ0,\operatorname{mass}B\lesssim\operatorname{mass}\beta+\epsilon\operatorname{mass}(\alpha)+\epsilon_{0},

so

FVZk(α)massβ\operatorname{FV}_{Z}^{k}(\alpha)\lesssim\operatorname{mass}\beta

as desired. ∎

The rest of this paper is dedicated to applying this theorem to horospheres and lattices in symmetric spaces and buildings.

3. Fillings in Sol2n1\operatorname{Sol}_{2n-1}

Theorem 1.3 is useful because it reduces a difficult-to-prove statement about the undistortedness of an inclusion to an easier-to-prove Lipschitz extension property. For example, in this section, we will prove:

Theorem 3.1.

The solvable Lie group Sol2n1=n1n\operatorname{Sol}_{2n-1}=\mathbb{R}^{n-1}\ltimes\mathbb{R}^{n} is Lipschitz (n2)(n-2)-connected.

Theorem 1.5 follows as a direct application of Theorem 1.3.

We start by defining Sol2n1\operatorname{Sol}_{2n-1}, n2n\geq 2. This group is a solvable Lie group which can be written as a semidirect product of n\mathbb{R}^{n} and n1\mathbb{R}^{n-1}, where n1\mathbb{R}^{n-1} acts on n\mathbb{R}^{n} as the group of n×nn\times n diagonal matrices with positive coefficients and determinant 1. When n=2n=2, this is the three-dimensional solvable group corresponding to solvegeometry.

All the constants and implicit constants in this section will depend on nn.

One feature of this group is that it can be realized as a horosphere in a product of hyperbolic planes. Let 2\mathbb{H}_{2} be the hyperbolic plane and let β:2\beta:\mathbb{H}_{2}\to\mathbb{R} be a Busemann function for 2\mathbb{H}_{2}. We can define Busemann functions β1,,βn\beta_{1},\dots,\beta_{n} in the product 2n\mathbb{H}_{2}^{n} by letting βi(x1,,xn)=β(xi)\beta_{i}(x_{1},\dots,x_{n})=\beta(x_{i}). Then b=n1/2βib=n^{-1/2}\sum\beta_{i} is a Busemann function for 2n\mathbb{H}_{2}^{n}, and Sol2n1\operatorname{Sol}_{2n-1} acts freely, isometrically, and transitively on the resulting horosphere b1(0)b^{-1}(0). The metric induced on Sol2n1\operatorname{Sol}_{2n-1} by inclusion in 2n\mathbb{H}_{2}^{n} is bilipschitz equivalent to a left-invariant Riemannian metric on Sol2n1\operatorname{Sol}_{2n-1}.

This group also appears as a subgroup of a Hilbert modular group. If ΓSL2()n\Gamma\subset\operatorname{SL}_{2}(\mathbb{R})^{n} is a Hilbert modular group and X=(2)nX=(\mathbb{H}_{2})^{n}, then there is a collection \mathcal{H} of disjoint open horoballs in XX such that the boundary of each horosphere is bilipschitz equivalent to Sol2n1\operatorname{Sol}_{2n-1} and Γ\Gamma acts cocompactly on XX\smallsetminus\mathcal{H} [Pit95]. Consequently, Theorem 1.6 is also a corollary of Theorem 3.1.

To prove Theorem 3.1, we will use the following condition, which is equivalent to Lipschitz connectivity (see [Gro96]):

Lemma 3.2.

Let ZZ be a metric space, let ΔZ\Delta_{Z} be the infinite-dimensional simplex with vertex set ZZ, and let ΔZ(k)\Delta_{Z}^{(k)} be its kk-skeleton. Let z0,,zk\langle z_{0},\dots,z_{k}\rangle denote the kk-simplex with vertices z0,,zkz_{0},\dots,z_{k}. Then ZZ is Lipschitz nn-connected if and only if there exists a map Ω:ΔZ(n+1)Z\Omega:\Delta_{Z}^{(n+1)}\to Z such that

  1. (1)

    For all zZz\in Z, Ω(z)=z\Omega(\langle z\rangle)=z.

  2. (2)

    There is a cc such that for any dn+1d\leq n+1 and any simplex δ=z0,,zd\delta=\langle z_{0},\dots,z_{d}\rangle, we have

    LipΩ|δcdiam{z0,,zd}.\operatorname{Lip}\Omega|_{\delta}\leq c\operatorname{diam}\{z_{0},\dots,z_{d}\}.
Proof.

One direction is clear; if ZZ is Lipschitz nn-connected, then one can construct Ω\Omega by letting Ω(z)=z\Omega(\langle z\rangle)=z for all zZz\in Z, then using the Lipschitz connectivity of ZZ to extend Ω\Omega over each skeleton inductively.

The other direction is an application of the Whitney decomposition. We view Dd+1D^{d+1} as a subset of d+1\mathbb{R}^{d+1}; by the Whitney covering lemma, the interior of Dd+1D^{d+1} can be decomposed into a union of countably many dyadic cubes such that for each cube CC, one has diamCdd(C,Sd)\operatorname{diam}C\sim_{d}d(C,S^{d}). We can decompose each cube into boundedly many simplices to get a triangulation τ\tau of the interior of Dd+1D^{d+1} where each simplex is bilipschitz equivalent to a scaling of the standard simplex.

We construct a map h:Dd+1Zh:D^{d+1}\to Z using this triangulation. For each vertex vv in τ\tau, let h(v)h(v) be a point in SdS^{d} such that d(v,h(v))d(v,h(v)) is minimized. One can check that hh is a Lipschitz map from τ(0)Sd\tau^{(0)}\to S^{d}, so g0=αh:τ(0)Zg_{0}=\alpha\circ h:\tau^{(0)}\to Z is a Lipschitz map with Lip(g0)d,ΩLip(α)\operatorname{Lip}(g_{0})\sim_{d,\Omega}\operatorname{Lip}(\alpha). We can extend g0g_{0} to a map g:τZg:\tau\to Z by sending the simplex v0,,vk\langle v_{0},\dots,v_{k}\rangle to the simplex Ω(g0(v0),,g0(vk))\Omega(\langle g_{0}(v_{0}),\dots,g_{0}(v_{k})\rangle), and this is also Lipschitz with Lip(g)d,ΩLip(α)\operatorname{Lip}(g)\sim_{d,\Omega}\operatorname{Lip}(\alpha).

Finally, we extend gg to a map β:Dd+1Z\beta:D^{d+1}\to Z by defining g(v)=α(v)g(v)=\alpha(v) when vSdv\in S^{d}. Since the diameter of the simplices of τ\tau goes to zero as one approaches the boundary, this extension is continuous and therefore Lipschitz, as desired. ∎

It therefore suffices to prove the following:

Lemma 3.3.

Let Δ=ΔSol2n1\Delta=\Delta_{\operatorname{Sol}_{2n-1}} be the infinite-dimensional simplex with vertex set Sol2n1\operatorname{Sol}_{2n-1}. There is a map Ω:Δ(n1)Sol2n1\Omega:\Delta^{(n-1)}\to\operatorname{Sol}_{2n-1} which satisfies the properties in Lemma 3.2. Therefore, Sol2n1\operatorname{Sol}_{2n-1} is Lipschitz (n2)(n-2)-connected.

Our construction is based on techniques from [BEW]; we will construct Ω\Omega using nonpositively curved subsets of Sol2n1\operatorname{Sol}_{2n-1} called kk-slices.

Recall that we defined Sol2n1\operatorname{Sol}_{2n-1} as a horosphere in (2)n(\mathbb{H}_{2})^{n}. Let β:2\beta:\mathbb{H}_{2}\to\mathbb{R} be the Busemann function used to define Sol2n1\operatorname{Sol}_{2n-1} and let * be the corresponding point at infinity. If γ\gamma is a geodesic in 2\mathbb{H}_{2} which has one endpoint at *, we call γ\gamma a vertical geodesic. For i=1,,ni=1,\dots,n, let si2s_{i}\subset\mathbb{H}_{2} be either a vertical geodesic or all of 2\mathbb{H}_{2}. If kk of the sis_{i}’s are equal to 2\mathbb{H}_{2}, we call the intersection s1××snSol2n1s_{1}\times\dots\times s_{n}\cap\operatorname{Sol}_{2n-1} a kk-slice.

Suppose that k<nk<n and that SS is a kk-slice; without loss of generality, we may assume that

S=2××2×γ1×γnkSol2n1.S=\mathbb{H}_{2}\times\dots\times\mathbb{H}_{2}\times\gamma_{1}\times\dots\gamma_{n-k}\cap\operatorname{Sol}_{2n-1}.

Then the projection to the first n1n-1 factors (i.e., all but the last factor) is a homeomorphism from SS to (2)k×nk1(\mathbb{H}_{2})^{k}\times\mathbb{R}^{n-k-1}. In fact, this map is bilipschitz, so SS is bilipschitz equivalent to a Hadamard manifold.

If k<nk<n, then any kk-slice is Lipschitz dd-connected for any dd:

Lemma 3.4.

If XX is a Hadamard manifold, it is Lipschitz nn-connected for any nn.

Proof.

Let α:SnX\alpha:S^{n}\to X, and let vSnv\in S^{n}. Let (x,r)Sn×[0,1](x,r)\in S^{n}\times[0,1] be polar coordinates on Dn+1D^{n+1}. We can construct a map α¯:Dn+1X\bar{\alpha}:D^{n+1}\to X by letting α¯(x,r)\bar{\alpha}(x,r) be the geodesic from α(v)\alpha(v) to α(x)\alpha(x). Because the distance function on XX is convex, this is a Lipschitz map with Lipschitz constant Lip(α)\lesssim\operatorname{Lip}(\alpha). ∎

If τ\tau is a polyhedral complex and f:τSol2n1f:\tau\to\operatorname{Sol}_{2n-1}, we say that ff is a slice map if the image of every cell δ\delta of τ\tau is contained in a (dimδ)(\dim\delta)-slice.

Our main tool in the proof of Lemma 3.3 is the following:

Lemma 3.5.

Let k<nk<n. Suppose that σ\sigma is a polyhedral complex which is bilipschitz equivalent to Sk1S^{k-1}. Then there is a c>0c>0 and a polyhedral complex τ\tau bilipschitz equivalent to DkD^{k} which has boundary σ\sigma. Furthermore, if f:σSol2n1f:\sigma\to\operatorname{Sol}_{2n-1} is a Lipschitz slice map, there is an extension g:τSol2n1g:\tau\to\operatorname{Sol}_{2n-1} which is a slice map with Lip(g)cLip(f)\operatorname{Lip}(g)\leq c\operatorname{Lip}(f).

The basic idea of the lemma is to first construct a family of projections along horospheres whose images lie in (n1)(n-1)-slices, then construct homotopies between ff and its projections. Gluing these homotopies together will give a map σ×[0,n]Sol2n1\sigma\times[0,n]\to\operatorname{Sol}_{2n-1}, and adding a final contraction will extend the map to all of τ\tau.

Let β:2\beta:\mathbb{H}_{2}\to\mathbb{R} be the Busemann function used to define Sol2n1\operatorname{Sol}_{2n-1}. If γ\gamma is a vertical geodesic in 2\mathbb{H}_{2} and x2x\in\mathbb{H}_{2}, let p(x)p(x) be the unique point on γ\gamma such that β(x)=β(p(x))\beta(x)=\beta(p(x)). This defines a map pγ:2γp_{\gamma}:\mathbb{H}_{2}\to\gamma. It is straightforward to check that pp is distance-decreasing and that d(x,p(x))2d(x,γ)d(x,p(x))\leq 2d(x,\gamma).

Suppose that 𝐱=(x1,,xn)(2)n\mathbf{x}=(x_{1},\dots,x_{n})\in(\mathbb{H}_{2})^{n}. For i=1,,ni=1,\dots,n, let γi\gamma_{i} be a vertical geodesic containing xix_{i}, and let β:2\beta:\mathbb{H}_{2}\to\mathbb{R} be the Busemann function used to define Sol2n1\operatorname{Sol}_{2n-1}. For each ii, let pi:Sol2n1nSol2n1p_{i}:\operatorname{Sol}_{2n-1}^{n}\to\operatorname{Sol}_{2n-1} be the map

pi(y1,,yn)=(y1,,yi1,pγi(yi),yi+1,,yn).p_{i}(y_{1},\dots,y_{n})=(y_{1},\dots,y_{i-1},p_{\gamma_{i}}(y_{i}),y_{i+1},\dots,y_{n}).

Let SS be the 0-slice

S=γ1××γnSol2n1S=\gamma_{1}\times\dots\times\gamma_{n}\cap\operatorname{Sol}_{2n-1}

and let SiS_{i} be the (n1)(n-1)-slice

Si=2××γi××2Sol2n1,S_{i}=\mathbb{H}_{2}\times\dots\times\gamma_{i}\times\dots\times\mathbb{H}_{2}\cap\operatorname{Sol}_{2n-1},

where γi\gamma_{i} occurs in the iith factor. It is easy to check the following properties:

  • pip_{i} is distance-decreasing

  • d(y,pi(y))2d(y,𝐱)d(y,p_{i}(y))\leq 2d(y,\mathbf{x}) for all ySol2n1y\in\operatorname{Sol}_{2n-1}

  • pip_{i} preserves SS pointwise

  • If SS^{\prime} is a dd-slice, then pi(S)p_{i}(S^{\prime}) lies in a dd-slice and SS^{\prime} and pi(S)p_{i}(S^{\prime}) both lie in the same (d+1)(d+1)-slice. In particular, yy and pi(y)p_{i}(y) lie in a 11-slice for every ySol2n1y\in\operatorname{Sol}_{2n-1}.

Then:

Lemma 3.6.

For any ii, if σ\sigma is a polyhedral complex with dimσ<n\dim\sigma<n, f:σSol2n1f:\sigma\to\operatorname{Sol}_{2n-1} is a Lipschitz slice map, and sσs\in\sigma satisfies f(s)=𝐱f(s)=\mathbf{x}, then there is a homotopy g:σ×[0,1]Sol2n1g:\sigma\times[0,1]\to\operatorname{Sol}_{2n-1} from ff to pifp_{i}\circ f which is a Lipschitz slice map with Lip(g)Lip(f)\operatorname{Lip}(g)\lesssim\operatorname{Lip}(f).

Proof.

We construct gg one skeleton at a time. For any cell δ\delta, the image g(δ×[0,1])g(\delta\times[0,1]) will be contained in the minimal slice that contains f(δ)f(\delta) and pi(f(δ))p_{i}(f(\delta)). Since f(δ)f(\delta) and pi(f(δ))p_{i}(f(\delta)) lie in a common (dimδ+1)(\dim\delta+1)-slice, this ensures that gg is a slice map.

The map gg is already defined on the vertices of σ×[0,1]\sigma\times[0,1] and we claim that it is Lipschitz on the 0-skeleton. If l=Lip(f)l=\operatorname{Lip}(f), then ff and pifp_{i}\circ f are ll-Lipschitz, and if vv is a vertex of σ\sigma, then

d(f(v),pi(f(v)))2d(f(v),𝐱)2ldiamσ,d(f(v),p_{i}(f(v)))\leq 2d(f(v),\mathbf{x})\leq 2l\operatorname{diam}\sigma,

so gg is Lipschitz on the vertex set with Lipschitz constant l\lesssim l.

Now suppose that we have defined gg on the (d1)(d-1)-cells of σ×[0,1]\sigma\times[0,1] and that for any (d2)(d-2)-cell δ\delta, the image g(δ×[0,1])g(\delta\times[0,1]) is contained in the minimal slice that contains f(δ)f(\delta) and pi(f(δ))p_{i}(f(\delta)). Consider a cell of the form δ×[0,1]\delta\times[0,1] for some (d1)(d-1)-cell δ\delta in σ\sigma. Since ff is a slice map, f(δ)f(\delta) lies in some (d1)(d-1)-slice, so f(δ)pi(f(δ))f(\delta)\cup p_{i}(f(\delta)) lies in some dd-slice, and this dd-slice also contains g(δ×[0,1])g(\partial\delta\times[0,1]) by the inductive hypothesis. Let SS^{\prime} be the minimal slice that contains

g((δ×[0,1]))=f(δ)pi(f(δ))g(δ×[0,1]).g(\partial(\delta\times[0,1]))=f(\delta)\cup p_{i}(f(\delta))\cup g(\partial\delta\times[0,1]).

By Lemma 3.4, we can extend gg over δ×[0,1]\delta\times[0,1] so that it sends δ×[0,1]\delta\times[0,1] to SS^{\prime}. The extension is Lipschitz and the Lipschitz constant is Lipf\lesssim\operatorname{Lip}f. ∎

Now we can prove Lemma 3.5.

Proof of Lemma 3.5.

Let τ\tau be the complex σ×[0,n]Cσ/\sigma\times[0,n]\cup C\sigma/\sim, where [0,n][0,n] is subdivided into nn unit-length edges, CσC\sigma is the cone over σ\sigma and \sim is the relation gluing the base of CσC\sigma to σ×{n}\sigma\times\{n\}. This is bilipschitz equivalent to DkD^{k}.

Choose a basepoint vσv_{*}\in\sigma and suppose that f(v)=𝐱=(x1,,xn)(2)nf(v_{*})=\mathbf{x}=(x_{1},\dots,x_{n})\in(\mathbb{H}_{2})^{n}. For i=0,,ni=0,\dots,n, let fi=pip1ff_{i}=p_{i}\circ\dots\circ p_{1}\circ f. By Lemma 3.6, for i=1,,ni=1,\dots,n, there is a homotopy gi:σ×[i1,i]Sol2n1g_{i}:\sigma\times[i-1,i]\to\operatorname{Sol}_{2n-1} from fi1f_{i-1} to fif_{i} which is a Lipschitz slice map with Lip(gi)Lip(f)\operatorname{Lip}(g_{i})\lesssim\operatorname{Lip}(f). Concatenating the gig_{i}’s gives a map σ×[0,n]Sol2n1\sigma\times[0,n]\to\operatorname{Sol}_{2n-1} which is a Lipschitz homotopy from ff to fnf_{n}. To define gg, it suffices to extend this map over CσC\sigma, but since the image of fnf_{n} lies in SS, we can use Lemma 3.4 to construct such an extension. Since this extension lies in a 0-slice, it is a slice map, so gg is a slice map and Lip(g)Lip(f)\operatorname{Lip}(g)\lesssim\operatorname{Lip}(f). ∎

Lemma 3.3 follows easily:

Proof of Lemma 3.3.

Let Δd\Delta_{d} be the standard dd-simplex. We define a sequence τi,i=0,,n1\tau_{i},i=0,\dots,n-1 of polyhedral complexes homeomorphic to Δi\Delta_{i} and a sequence σi,i=0,,n1\sigma_{i},i=0,\dots,n-1 of polyhedral complexes homeomorphic to Δi+1\partial\Delta_{i+1} inductively. Let τ0\tau_{0} be a single point. For each i0i\geq 0, let σi\sigma_{i} be the complex obtained by replacing each ii-face of Δi+1\partial\Delta_{i+1} by a copy of τi\tau_{i}. Let τi+1\tau_{i+1} be the complex obtained by applying Lemma 3.5 to σi\sigma_{i}. This is PL-homeomorphic to Δi\Delta_{i} and has boundary σi\sigma_{i}.

Let Δ\Delta^{\prime} be the complex obtained by subdividing each dd-simplex of Δ(n1)\Delta^{(n-1)} into a copy of τd\tau_{d} and let i:Δ(n1)Δi:\Delta^{(n-1)}\to\Delta^{\prime} be a bilipschitz equivalence taking each simplex to the corresponding copy of τd\tau_{d}. We can construct a slice map Ω:ΔZ\Omega^{\prime}:\Delta^{\prime}\to Z by defining Ω(x)=x\Omega^{\prime}(\langle x\rangle)=x for all xSol2n1x\in\operatorname{Sol}_{2n-1} and using Lemma 3.5 inductively to extend Ω\Omega^{\prime} over each of the τd\tau_{d}’s.

That is, if δ=x0,,xd+1Δ\delta=\langle x_{0},\dots,x_{d+1}\rangle\subset\Delta is a (d+1)(d+1)-cell, Ω\Omega^{\prime} is defined on i(δ)i(\partial\delta), and

Ω|i(δ):σdSol2n1\Omega|_{i(\partial\delta)}:\sigma_{d}\to\operatorname{Sol}_{2n-1}

is a slice map with Lipschitz constant diam{x0,,xd+1}\lesssim\operatorname{diam}\{x_{0},\dots,x_{d+1}\}, we may extend it to a slice map on i(δ)i(\delta) using Lemma 3.5. The resulting map Ω|δ\Omega|_{\delta} has

Lip(Ω|δ)diam{x0,,xd+1}\operatorname{Lip}(\Omega|_{\delta})\lesssim\operatorname{diam}\{x_{0},\dots,x_{d+1}\}

as desired. ∎

Thus, by Lemma 3.2, Sol2n1\operatorname{Sol}_{2n-1} is Lipschitz (n2)(n-2)-connected, and by Theorem 1.3, it is undistorted up to dimension nn inside 2n\mathbb{H}_{2}^{n}. Consequently, if k<nk<n and if α\alpha is a Lipschitz kk-cycle in Sol2n1\operatorname{Sol}_{2n-1}, then

FVSol2n1k+1(α)FV2nk+1(α)(massα)(k+1)/k,\operatorname{FV}_{\operatorname{Sol}_{2n-1}}^{k+1}(\alpha)\lesssim\operatorname{FV}_{\mathbb{H}_{2}^{n}}^{k+1}(\alpha)\lesssim(\operatorname{mass}\alpha)^{(k+1)/k},

as desired.

4. Fillings in horospheres of euclidean buildings

In this section, we prove Theorem 1.7.

We claim:

Theorem 4.1.

Let XX be a thick euclidean building and let XX_{\infty} be the Bruhat-Tits building of XX. If XX is reducible, then XX_{\infty} is a join of buildings; let τ\tau be a point in XX_{\infty} which has rational slope and is not contained in a join factor of rank less than nn. Let ZZ be a horosphere in XX centered at τ\tau and let p:XM(X)p:X_{\infty}\to M(X_{\infty}) be the projection of XX_{\infty} to its model chamber. Then ZZ is Lipschitz (n2)(n-2)-connected with implicit constant depending only on XX and p(τ)p(\tau).

By Theorem 1.3 and Corollary 1.4, this implies Theorem 1.7.

Furthermore, if KK is a global function field, 𝐆\mathbf{G} is a noncommutative, absolutely almost simple KK-group of KK-rank 1, and SS is a finite set of pairwise inequivalent valuations on KK, then Γ=𝐆(𝒪S)\Gamma=\mathbf{G}(\mathcal{O}_{S}) is an SS-arithmetic group. If XX is the associated euclidean building and nn is its rank, then by Theorem 3.7 of [BW11], there is a collection \mathcal{H} of pairwise disjoint open horoballs in XX such that XX\smallsetminus\mathcal{H} is 𝐆(𝒪S)\mathbf{G}(\mathcal{O}_{S})-invariant and cocompact. By Theorem 4.1, the boundary of each of these horoballs is Lipschitz (n2)(n-2)-connected with a uniform implicit constant, so Theorem 1.3 implies Theorem 1.8.

As in [Dru04, Rem. 4.2], it suffices to consider the case that XX is a thick euclidean building of rank nn and that τ\tau is not parallel to any factor of XX. If X=X1×X2X=X_{1}\times X_{2}, then X=(X1)(X2)X_{\infty}=(X_{1})_{\infty}\ast(X_{2})_{\infty}. If τ(X1)\tau\in(X_{1})_{\infty}, then Z=Z1×X2Z=Z_{1}\times X_{2}, where Z1X1Z_{1}\subset X_{1} is a horosphere of X1X_{1} centered at τ\tau. If α:SkZ\alpha:S^{k}\to Z is cc-Lipschitz, we can replace it with its projection to Z1Z_{1} by using an homotopy with Lipschitz constant c\lesssim c, so if Z1Z_{1} is Lipschitz (n2)(n-2)-connected, so is ZZ.

Therefore, in this section, we will let XX be a thick euclidean building of rank nn equipped with its complete apartment system, and let XX_{\infty} be its Bruhat-Tits building. We fix a direction at infinity τX\tau\in X_{\infty} which is not contained in any factor of XX_{\infty}, and let hh be a Busemann function centered at τ\tau, with corresponding horosphere Z=h1(0)Z=h^{-1}(0). We orient hh so that h(x)h(x) increases as xx approaches τ\tau; we use this orientation so that we can treat hh as a Morse function on XX more easily.

All the constants in this section and its subsections will depend on XX and ZZ.

The proof that ZZ is Lipschitz (n2)(n-2)-connected is based on Lemma 3.2. Let ΔZ\Delta_{Z} be the infinite-dimensional simplex with vertex set ZZ, and let ΔZ(k)\Delta_{Z}^{(k)} be its kk-skeleton. We will show:

Lemma 4.2.

There exists a map Ω:ΔZ(n1)Z\Omega:\Delta_{Z}^{(n-1)}\to Z such that

  1. (1)

    For all zZz\in Z, Ω(z)=z\Omega(\langle z\rangle)=z.

  2. (2)

    For any dn+1d\leq n+1 and any simplex δ=z0,,zd\delta=\langle z_{0},\dots,z_{d}\rangle, we have

    LipΩ|δdiam{z0,,zd}+1.\operatorname{Lip}\Omega|_{\delta}\lesssim\operatorname{diam}\{z_{0},\dots,z_{d}\}+1.

The only difference between the map in Lemma 4.2 and the map in Lemma 3.2 is the bound on LipΩ|δ\operatorname{Lip}\Omega|_{\delta}. In Lemma 3.2, LipΩ|δ\operatorname{Lip}\Omega|_{\delta} is bounded by a multiple of diam{z0,,zd}\operatorname{diam}\{z_{0},\dots,z_{d}\}; in Lemma 4.2, it is bounded by a multiple of diam{z0,,zd}\operatorname{diam}\{z_{0},\dots,z_{d}\} and an additive constant.

As a corollary, we have:

Lemma 4.3.

For any t>0t>0, there is a Lipschitz map rt:h1((,t])X(n1)Zr_{t}:h^{-1}((\infty,t])\cap X^{(n-1)}\to Z which restricts to the identity map on ZZ.

Proof.

Define rtr_{t} on h1((,0])h^{-1}((\infty,0]) as the closest-point projection. Since horoballs are convex, this is a distance-decreasing map.

To define rtr_{t} on h1((0,t])X(n1)h^{-1}((0,t])\cap X^{(n-1)}, we view XX as a polyhedral complex, i.e., a complex whose faces consist of convex polyhedra in n\mathbb{R}^{n}, glued along faces by isometries. Then hh is linear on each face of XX, so if PP is a face of XX, then the intersections h1([0,t])Ph^{-1}([0,t])\cap P, ZPZ\cap P, and h1(t)Ph^{-1}(t)\cap P are convex polyhedra. Since τ\tau has rational slope, the set h(X(0))h(X^{(0)}) of possible values of hh on the vertices of XX is discrete, so only finitely many isometry classes of polyhedra occur this way, and we can give Zt=h1([0,t])X(n1)Z_{t}=h^{-1}([0,t])\cap X^{(n-1)} the structure of a polyhedral complex with only finitely many isometry classes of cells. We subdivide each cell to make ZtZ_{t} into a simplicial complex. We define rtr_{t} on the vertices of ZtZ_{t} so that d(rt(v),v)d(r_{t}(v),v) is minimized. If Δ\Delta is a simplex of ZtZ_{t} with vertices v0,,vkv_{0},\dots,v_{k}, we define

rt|Δ=Ω|rt(v0),,rt(vk).r_{t}|_{\Delta}=\Omega|_{\rangle r_{t}(v_{0}),\dots,r_{t}(v_{k})}\rangle.

This gives a Lipschitz map with Lipschitz constant depending on the size of the smallest simplex in ZtZ_{t}. ∎

The proof of this lemma is the only place that we use the assumption that τ\tau has rational slope.

Given these lemmas, we prove Theorem 4.1 as follows:

Proof of Theorem 4.1.

Suppose that α:SkZ\alpha:S^{k}\to Z is a Lipschitz map. If Lip(α)1\operatorname{Lip}(\alpha)\leq 1, we can extend α\alpha to a map β:Dk+1X\beta:D^{k+1}\to X by coning α\alpha to a point along geodesics in XX. Since XX is CAT(0), β\beta is Lipschitz with Lipschitz constant Lip(α)\sim\operatorname{Lip}(\alpha). Furthermore, the image of β\beta lies in h1([1,1])h^{-1}([-1,1]), so r1β:Dk+1Zr_{1}\circ\beta:D^{k+1}\to Z is an extension of α\alpha with Lip(r1β)Lip(α)\operatorname{Lip}(r_{1}\circ\beta)\sim\operatorname{Lip}(\alpha).

If Lip(α)>1\operatorname{Lip}(\alpha)>1, let LL\in\mathbb{N} be the smallest integer such that LLip(α)L\geq\operatorname{Lip}(\alpha), let Dk+1(L)D^{k+1}(L) be the cube [0,L]k+1k+1[0,L]^{k+1}\subset\mathbb{R}^{k+1}, and let Sk(L)=Dk(L)S^{k}(L)=\partial D^{k}(L). We view α\alpha as a map Sk(L)ZS^{k}(L)\to Z with Lipschitz constant 1\sim 1 and try to construct an extension to Dk+1(L)D^{k+1}(L) with Lipschitz constant 1\sim 1.

As in the proof of Lemma 3.2, the Whitney covering lemma implies that Dk+1(L)D^{k+1}(L) can be decomposed into a union of countably many dyadic cubes such that for each cube CC, one has diamCd(C,Sk(L))\operatorname{diam}C\sim d(C,S^{k}(L)). Since these cubes are dyadic, each cube of side length less than one is contained in a cube of side length 1. Let 𝒞\mathcal{C} be the cover of Dk+1(L)D^{k+1}(L) obtained by combining cubes of side length less than 1 into cubes of side length 1. Then for each cube CC in 𝒞\mathcal{C}, we have diamCd(C,Sk(L))+1\operatorname{diam}C\sim d(C,S^{k}(L))+1, and each cube which touches Sk(L)S^{k}(L) has side length 1. We call the cubes that touch Sk(L)S^{k}(L) the boundary cubes and we call the rest interior cubes. We can decompose each cube into boundedly many simplices to get a triangulation τ\tau of Dd+1D^{d+1} where each simplex is bilipschitz equivalent to a scaling of the standard simplex. Let τi\tau_{i} be the subcomplex of τ\tau contained in the interior cubes.

We construct a map h:Sk(L)τiZh:S^{k}(L)\cup\tau_{i}\to Z using this triangulation. If xSk(L)x\in S^{k}(L), we define h(x)=f(x)h(x)=f(x). For each vertex vv in τi\tau_{i}, let h(v)h(v) be a point in SdS^{d} such that d(v,h(v))d(v,h(v)) is minimized, and if Δ=v0,,vk\Delta=\langle v_{0},\dots,v_{k}\rangle is a simplex of τi\tau_{i}, define

h|Δ=Ω|h(v0),,h(vk).h|_{\Delta}=\Omega|_{\langle h(v_{0}),\dots,h(v_{k})\rangle}.

Since diamΔ1\operatorname{diam}\Delta\gtrsim 1, this is Lipschitz with Lip(h)1\operatorname{Lip}(h)\sim 1.

Since XX is CAT(0) and thus Lipschitz nn-connected, we can extend hh over the boundary cubes inductively; if CC is a face of a boundary cube and hh is already defined on C\partial C, we extend hh over CC by coning h|Ch|_{\partial C} to a point along geodesics. This produces an extension h:Dk+1(L)Xh:D^{k+1}(L)\to X with Lipschitz constant Lip(h)1\operatorname{Lip}(h)\sim 1.

Finally, since the boundary cubes are all contained in a neighborhood of Sk(L)S^{k}(L), their image is contained in a neighborhood of ZZ, so if cc is large enough, then rch:Dk(L)Zr_{c}\circ h:D^{k}(L)\to Z is an extension of α\alpha with Lipschitz constant 1\sim 1. ∎

In the rest of this section, we will prove Lemma 4.2. The proof is a quantitative Morse theory argument, like the “pushing” arguments in [ABD+12]. Bux and Wortman [BW11] used a Morse theory argument to prove that ZZ is nn-connected; we sketch their proof in the case that τ\tau is a generic direction. In general, XX is contractible, and ZZ is the level set of hh. If τ\tau is generic, then hh is nonconstant on every edge of XX, and we can treat it as a combinatorial Morse function.

That is, if uu is a vertex of XX, then every vertex of its link Lk(u)\operatorname{Lk}(u) corresponds to a vertex vv adjacent to uu. We define the downward link LkuLku\operatorname{\text{Lk}^{\downarrow}}{u}\subset\operatorname{Lk}u to be the subcomplex spanned by vertices vv with h(v)<h(u)h(v)<h(u). By results of Schulz [Sch], Lku\operatorname{\text{Lk}^{\downarrow}}{u} is (n2)(n-2)-connected for all uu, so combinatorial Morse theory implies that ZZ is also (n2)(n-2)-connected. Bux and Wortman apply a similar argument in the general case, but with hh replaced by a more complicated function to deal with faces of dimension >0>0 that are orthogonal to τ\tau.

Arguments like this, however, give poor quantitative bounds. Given an (n2)(n-2)-sphere in ZZ, one can construct a filling in the horosphere h1([0,))h^{-1}([0,\infty)) and use Morse theory to homotope it to ZZ, but the filling may grow exponentially large in the process. The pushing methods in [ABD+12] avoid this sort of exponential growth by constructing maps from Lku\operatorname{\text{Lk}^{\downarrow}}{u} to ZZ, and we will apply similar methods here.

Let 𝔞\mathfrak{a} be a chamber of XX_{\infty} which contains τ\tau in its closure and let

X0(𝔞):={𝔟𝔟 is a chamber of X opposite to 𝔞}.X_{\infty}^{0}(\mathfrak{a}):=\{\mathfrak{b}\mid\mathfrak{b}\text{ is a chamber of $X_{\infty}$ opposite to $\mathfrak{a}$}\}.

Abramenko [Abr96] showed that if YY is a sufficiently thick classical spherical building, then Y0(𝔞)Y^{0}(\mathfrak{a}) is (rkY2)(\operatorname{rk}Y-2)-connected for any chamber 𝔞\mathfrak{a} of YY. We will show that if XX is a thick euclidean building of rank nn, then X0(𝔞)X_{\infty}^{0}(\mathfrak{a}) is (n2)(n-2)-connected.

Roughly, we show (n2)(n-2)-connectivity by showing that “most” pairs of chambers 𝔟,𝔠X0(𝔞)\mathfrak{b},\mathfrak{c}\subset{X_{\infty}^{0}(\mathfrak{a})} are opposite to one another and that if E𝔟,𝔠E_{\mathfrak{b},\mathfrak{c}} is the apartment they span, then E𝔟,𝔠X0(𝔞)\partial_{\infty}E_{\mathfrak{b},\mathfrak{c}}\subset{X_{\infty}^{0}(\mathfrak{a})}. Then, for each sphere α:SkX0(𝔞)\alpha:S^{k}\to{X_{\infty}^{0}(\mathfrak{a})} with k<n2k<n-2, we choose a 𝔠\mathfrak{c} such that for any 𝔟\mathfrak{b} in the support of α(Sn2)\alpha(S^{n-2}), 𝔟\mathfrak{b} is opposite to 𝔠\mathfrak{c} and E𝔟,𝔠X0(𝔞)E_{\mathfrak{b},\mathfrak{c}}\subset{X_{\infty}^{0}(\mathfrak{a})}. We can then contract α\alpha to a point in 𝔠\mathfrak{c} along geodesics. Since X0(𝔞)X_{\infty}^{0}(\mathfrak{a}) is (n2)(n-2)-connected, there is no obstruction to constructing a map

Ω:ΔZ(n1)X0(𝔞).\Omega_{\infty}:\Delta_{Z}^{(n-1)}\to X_{\infty}^{0}(\mathfrak{a}).

Next, we construct a map to ZZ. Given a point xXx\in X and a direction σX\sigma\in X_{\infty}, we let rr be the ray emanating from xx in the direction of σ\sigma. If h(x)>0h(x)>0 and σX0(𝔞)\sigma\in{X_{\infty}^{0}(\mathfrak{a})}, this ray will eventually intersect ZZ. This provides a map X0(𝔞)ZX_{\infty}^{0}(\mathfrak{a})\to Z, but this map is not Lipschitz – a ray may travel a long distance before intersecting ZZ. To fix this, we define the downward link at infinity Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) at xx. This is a subset Lk(x)X\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)\subset X_{\infty} of directions that point “downward” from xx (i.e., away from 𝔞\mathfrak{a}). Rays in these directions intersect ZZ after traveling distance h(x)\lesssim h(x), so we can define a map ix:Lk(x)Zi_{x}:\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)\to Z with Lipschitz constant h(x)\lesssim h(x) which sends each direction to the corresponding intersection point.

The sets Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) get bigger as x𝔞x\to\mathfrak{a}, and any finite subset of X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} is contained in some Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x). This lets us convert Ω\Omega_{\infty} into a map to ZZ; for each simplex Δ\Delta, we choose some xΔx_{\Delta} so that Ω(Δ)Lk(xΔ)\Omega_{\infty}(\Delta)\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x_{\Delta}) and define (after some patching around the edges)

Ω|Δ=ixΔΩ.\Omega|_{\Delta}=i_{x_{\Delta}}\circ\Omega_{\infty}.

Finally, we show that restrictions of Ω\Omega to simplices satisfy Lipschitz bounds. To do this, we need some control over the Lipschitz constants of the ixΔi_{x_{\Delta}}’s. The Lipschitz constant of ixΔi_{x_{\Delta}} is on the order of h(xΔ)h(x_{\Delta}), so we try to bound the h(xΔ)h(x_{\Delta})’s by controlling which chambers of X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} we use in fillings of spheres. This proves the theorem.

The rest of this section is devoted to filling in the details of this sketch. First, in Sections 4.1 and 4.2, we describe our notation and define some maps and subsets that we will use in the rest of the proof. In Section 4.3, we construct Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) and show that there are many apartments in Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) . In Section 4.4, we use this fact to show that X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} is (n2)(n-2)-connected and to construct Ω\Omega_{\infty} and the xΔx_{\Delta}’s. In Section 4.5, we use these to construct Ω\Omega.

4.1. Preliminaries

In this section, we fix some notation for dealing with buildings, define some maps and subsets that will be important in the rest of the section, and prove some of their properties. Our primary reference is [AB08].

As stated in the introduction to this section, we let XX be an irreducible thick euclidean building of rank nn, equipped with its complete apartment system and let XX_{\infty} be its Bruhat-Tits building. If EE is an apartment of XX, we can identify it with the Coxeter complex of a Euclidean reflection group WW, and we can identify the corresponding apartment EX\partial_{\infty}E\subset X_{\infty} with the Coxeter complex of W¯\bar{W}, the reflection group corresponding to the linear parts of the elements of WW.

Recall that XX_{\infty} can be defined as the set of classes of parallel unit-speed geodesic rays in XX, where r,r:[0,)Xr,r^{\prime}:[0,\infty)\to X are parallel if d(r(t),r(t))d(r(t),r^{\prime}(t)) is bounded as tt\to\infty. For any xXx\in X and any σX\sigma\in X_{\infty}, there is a unique ray based at xx and parallel to σ\sigma [AB08, Lem. 11.72]. Given a subset YXY\subset X, we define Y\partial_{\infty}Y to be its boundary at infinity; for the subsets we will consider in this paper, Y\partial_{\infty}Y consists of the set of parallelism classes of geodesic rays in YY. If 𝔡\mathfrak{d} is a chamber of E\partial_{\infty}E, we say that EE is asymptotic to 𝔡\mathfrak{d}.

If xEx\in E, there is a conical cell x+𝔡x+\mathfrak{d} based at xx for every chamber 𝔡\mathfrak{d} of E\partial_{\infty}E; we call these cells sectors. Note that x+𝔡x+\mathfrak{d} doesn’t depend on our choice of EE; this construction gives the same result for any apartment EE^{\prime} such that 𝔡E\mathfrak{d}\subset\partial_{\infty}E^{\prime} and xEx\in E^{\prime}.

The codimension-1 cells of EE are called panels. Each panel is contained in a codimension-1 subspace of EE which we call a wall. Each wall divides EE into a pair of closed half-apartments. We say that EE^{\prime} is a ramification of EE if either E=EE=E^{\prime} or EEE\cap E^{\prime} is a half-apartment. Since XX is thick, each wall is the boundary of at least three half-apartments. We say that two chambers are adjacent if they have disjoint interiors and share a panel. A sequence of chambers C1,,CkC_{1},\dots,C_{k} such that CiC_{i} and Ci+1C_{i+1} are adjacent is called a gallery of combinatorial length kk. The minimal combinatorial length of a gallery connecting two chambers is called the combinatorial distance between them, and a gallery realizing this length is called a minimal gallery. We denote the combinatorial distance between CC and CC^{\prime} by dcomb(C,C)d_{\text{comb}}(C,C^{\prime}).

There is also a CAT(0) metric on XX which gives each apartment the metric of n\mathbb{R}^{n}. We denote this metric by d:X×Xd:X\times X\to\mathbb{R}. Likewise, there is a CAT(1) metric (the angular metric) on XX_{\infty}, which we also denote by dd.

If 𝔠,𝔡X\mathfrak{c},\mathfrak{d}\subset X_{\infty} are chambers and dcomb(𝔠,𝔡)=diamcomb(X)d_{\text{comb}}(\mathfrak{c},\mathfrak{d})=\operatorname{diam}_{\text{comb}}(X_{\infty}), we say that 𝔠\mathfrak{c} and 𝔡\mathfrak{d} are opposite. Any pair of opposite chambers of XX_{\infty} determines a unique apartment of XX_{\infty} [AB08, Thm. 4.70]. Indeed, if 𝔠,𝔡X\mathfrak{c},\mathfrak{d}\subset X_{\infty} are opposite chambers, then there is a unique apartment of XX which is asymptotic to 𝔠\mathfrak{c} and 𝔡\mathfrak{d} [AB08, Thm. 11.63].

4.2. Folded apartments

In order to prove Theorem 4.1, we will need to understand how apartments in XX are positioned relative to 𝔞\mathfrak{a}. In this section, we describe some notions that will be useful to understand the arrangement of apartments in XX.

Recall that if EE is an apartment of XX and CEC\subset E is a chamber, there is a retraction ρE,C:XE\rho_{E,C}:X\to E such that if C=C1,,CkC=C_{1},\dots,C_{k} is a minimal gallery in XX, then C=ρ(C1),,ρ(Ck)C=\rho(C_{1}),\dots,\rho(C_{k}) is a minimal gallery in EE. We will use a related retraction which is based at a chamber of XX_{\infty} rather than a chamber of XX.

Following Abramenko and Brown [AB08, 11.7], if EE is an apartment of XX and 𝔠\mathfrak{c} is a chamber of E\partial_{\infty}E, we define ρE,𝔠:XE\rho_{E,\mathfrak{c}}:X\to E to be the map such that if EE^{\prime} is an apartment of XX which is asymptotic to 𝔠\mathfrak{c}, then ρE,𝔠|E\rho_{E,\mathfrak{c}}|_{E^{\prime}} is the isomorphism ϕE:EE\phi_{E^{\prime}}:E^{\prime}\to E which fixes EEE\cap E^{\prime} pointwise. (In the case that XX is a tree, this is the map obtained by “dangling” the tree from a point at infinity.)

Fix some apartment FF which is asymptotic to 𝔞\mathfrak{a} and let ρ=ρF,𝔞\rho=\rho_{F,\mathfrak{a}}. Note that changing the choice of FF changes ρ\rho by an isomorphism; if FF^{\prime} is asymptotic to 𝔞\mathfrak{a} and ϕF:FF\phi_{F}:F\to F^{\prime} is the isomorphism fixing FFF\cap F^{\prime} pointwise, then ρF,𝔞=ϕFρF,𝔞\rho_{F^{\prime},\mathfrak{a}}=\phi_{F}\circ\rho_{F,\mathfrak{a}}. Furthermore, ρ\rho preserves Busemann functions centered at points in 𝔞\mathfrak{a}. In particular, hρ=hh\circ\rho=h.

If EE is an apartment of XX, then ρ\rho maps EE to FF by a “folding” process. If XX is a tree, for instance, then either ρ|E\rho|_{E} is an isomorphism EFE\to F or it folds EE once. In higher rank buildings, ρ|E\rho|_{E} can be more complicated. The following lemmas will help us describe these maps.

For any chamber CC of XX and any chamber 𝔠\mathfrak{c} of XX_{\infty}, we define the direction DC(𝔠)D_{C}(\mathfrak{c}) of ρ(𝔠)\rho(\mathfrak{c}) at ρ(C)\rho(C) as follows. Let xy\overrightarrow{xy} be a directed line segment in CC in the direction of an interior point of 𝔠\mathfrak{c}. Then ρ(xy)\rho(\overrightarrow{xy}) is a directed line segment in FF pointing toward the interior of some chamber of F\partial_{\infty}F. We let DC(𝔠)D_{C}(\mathfrak{c}) be that chamber.

Lemma 4.4.

Let CC be a chamber of an apartment EE. Then DC|E:EFD_{C}|_{\partial_{\infty}E}:\partial_{\infty}E\to\partial_{\infty}F is a type-preserving isomorphism.

Proof.

If EE^{\prime} is an apartment containing CC and asymptotic to 𝔞\mathfrak{a} and 𝔠E\mathfrak{c}^{\prime}\subset\partial_{\infty}E^{\prime}, we have DC(𝔠)=ρ(𝔠)D_{C}(\mathfrak{c}^{\prime})=\rho_{\infty}(\mathfrak{c}^{\prime}). If ϕ:EE\phi:E\to E^{\prime} is the isomorphism fixing EEE\cap E^{\prime} pointwise, then DC(𝔠)=DC(ϕ(𝔠))D_{C}(\mathfrak{c})=D_{C}(\phi_{\infty}(\mathfrak{c})) for any 𝔠E\mathfrak{c}\subset\partial_{\infty}E, so

DC|E=ρ|Eϕ.D_{C}|_{\partial_{\infty}E}=\rho_{\infty}|_{\partial_{\infty}E^{\prime}}\circ\phi_{\infty}.

By Proposition 11.87 of [AB08], ϕ\phi_{\infty} is a type-preserving isomorphism. Likewise, since ρ|E\rho|_{E^{\prime}} is the isomorphism fixing EFE^{\prime}\cap F pointwise, it induces a type-preserving isomorphism on E\partial_{\infty}E^{\prime}. ∎

Refer to caption𝔞\mathfrak{a}
Figure 1. A subset of an apartment and its image under ρ\rho. (The three-dimensional effect is for clarity – the map sends triangles to triangles.) Each triangle is 𝔞\mathfrak{a}-characteristic for the chamber of XX_{\infty} in the direction of its arrow.

If CC is a chamber of XX, xCx\in C, and 𝔠X\mathfrak{c}\subset X_{\infty}, then there is some subsector x+𝔠x^{\prime}+\mathfrak{c} of x+𝔠x+\mathfrak{c} such that some apartment of XX contains x+𝔠x^{\prime}+\mathfrak{c} and is asymptotic to 𝔞\mathfrak{a}. The proof of Theorem 11.63 (2) in [AB08] contains the following lemma, which gives us a criterion for when we can take x=xx^{\prime}=x.

Lemma 4.5.

Suppose that EE is an apartment of XX and 𝔠\mathfrak{c} is a chamber in E\partial_{\infty}E. If CC is a chamber of EE such that

dcomb(𝔞,DC(𝔠))=maxBEdcomb(𝔞,DB(𝔠))d_{\text{comb}}(\mathfrak{a},D_{C}(\mathfrak{c}))=\max_{B\subset E}d_{\text{comb}}(\mathfrak{a},D_{B}(\mathfrak{c}))

and xCx\in C, then there is an apartment of XX containing x+𝔠x+\mathfrak{c} and asymptotic to 𝔞\mathfrak{a}.

In particular, if 𝔞\mathfrak{a} and DC(𝔠)D_{C}(\mathfrak{c}) are opposite, then 𝔞\mathfrak{a} and 𝔠\mathfrak{c} are opposite.

If CC is a chamber of XX and 𝔠\mathfrak{c} is a chamber of X\partial_{\infty}X such that 𝔞\mathfrak{a} is opposite to DC(𝔠)D_{C}(\mathfrak{c}), we call CC an 𝔞\mathfrak{a}-characteristic chamber for 𝔠\mathfrak{c}.

Lemma 4.6.

The following are equivalent:

  • CC is an 𝔞\mathfrak{a}-characteristic chamber for 𝔠\mathfrak{c}.

  • 𝔞\mathfrak{a} and 𝔠\mathfrak{c} are opposite and the unique apartment asymptotic to 𝔞\mathfrak{a} and 𝔠\mathfrak{c} contains CC.

  • 𝔞\mathfrak{a} and 𝔠\mathfrak{c} point in opposite directions at CC. That is, whenever xx is in the interior of CC, the rays from xx toward the barycenters of 𝔞\mathfrak{a} and 𝔠\mathfrak{c} point in opposite directions.

Proof.

(1) implies (2) by Lemma 4.5. If (2) holds and EE is the unique apartment asymptotic to 𝔞\mathfrak{a} and 𝔠\mathfrak{c}, then the rays toward the barycenters of 𝔞\mathfrak{a} and 𝔠\mathfrak{c} from any point in EE are rays in EE pointing in opposite directions, so (3) holds. Finally, if (3) holds, then DC(𝔞)D_{C}(\mathfrak{a}) and DC(𝔠)D_{C}(\mathfrak{c}) are opposite chambers of F\partial_{\infty}F. Since DC(𝔞)=𝔞D_{C}(\mathfrak{a})=\mathfrak{a}, this implies (1). ∎

We can replace 𝔞\mathfrak{a} in the above constructions with any chamber 𝔡X\mathfrak{d}\subset X_{\infty}, so more generally, we may say that CC is an 𝔡\mathfrak{d}-characteristic chamber for 𝔠\mathfrak{c} if 𝔡\mathfrak{d} and 𝔠\mathfrak{c} are opposite and the unique apartment asymptotic to 𝔡\mathfrak{d} and 𝔠\mathfrak{c} contains CC. Then CC is an 𝔡\mathfrak{d}-characteristic chamber for 𝔠\mathfrak{c} if and only if 𝔞\mathfrak{a} and 𝔠\mathfrak{c} point in opposite directions at CC.

Similarly, we say that 𝔠\mathfrak{c} and 𝔠\mathfrak{c}^{\prime} point in the same direction at CC if, whenever xx is in the interior of CC, the rays from xx toward the barycenters of 𝔠\mathfrak{c} and 𝔠\mathfrak{c}^{\prime} have the same tangent vector at xx. It follows that

Lemma 4.7.

If 𝔠\mathfrak{c} and 𝔠\mathfrak{c}^{\prime} point in the same direction at CC and CC is 𝔡\mathfrak{d}-characteristic for 𝔠\mathfrak{c}, then it is also 𝔡\mathfrak{d}-characteristic for 𝔠\mathfrak{c}^{\prime}.

We can apply this lemma to ramifications: if CEC\subset E is 𝔞\mathfrak{a}-characteristic for 𝔠E\mathfrak{c}\subset\partial_{\infty}E and EE^{\prime} is any apartment of XX that contains CC, let ϕ:EE\phi:E\to E^{\prime} be the isomorphism fixing EEE\cap E^{\prime} pointwise and let 𝔠=ϕ(𝔠)\mathfrak{c}^{\prime}=\phi_{\infty}(\mathfrak{c}). Then 𝔠\mathfrak{c} and 𝔠\mathfrak{c}^{\prime} point in the same direction at CC, so 𝔠\mathfrak{c}^{\prime} is opposite to 𝔞\mathfrak{a}.

Figure 1 gives an example of the possible behavior of ρ\rho on an apartment; in the figure, ρ\rho “folds” EE along the thick lines. Each of the arrows is sent to an arrow pointing in the direction opposite 𝔞\mathfrak{a}, so each chamber of EE is 𝔞\mathfrak{a}-characteristic for the chamber of EE that its arrow points toward. Since there are arrows pointing toward every chamber of E\partial_{\infty}E, we have EX0(𝔞)\partial_{\infty}E\subset{X_{\infty}^{0}(\mathfrak{a})}. Any apartment EE^{\prime} that contains the pictured portion of EE also satisfies EX0(𝔞)\partial_{\infty}E^{\prime}\subset{X_{\infty}^{0}(\mathfrak{a})}. In fact, if EE^{\prime} is such an apartment, then ρ\rho “folds” EE^{\prime} in the same way as EE (i.e., if ϕ:EE\phi:E\to E^{\prime} is the isomorphism fixing EEE\cap E^{\prime} pointwise, then ρ|E=ρ|Eϕ\rho|_{E}=\rho|_{E^{\prime}}\circ\phi).

As the figure suggests, every apartment can be decomposed into 𝔞\mathfrak{a}-characteristic chambers:

Lemma 4.8 (see [Dru04, Lem. 3.1.1]).

If EE is an apartment of XX and 𝔠1,,𝔠dE\mathfrak{c}_{1},\dots,\mathfrak{c}_{d}\in\partial_{\infty}E are the chambers of E\partial_{\infty}E which are opposite to 𝔞\mathfrak{a}, then EE is a union of subcomplexes Y1,,YdY_{1},\dots,Y_{d} such that the chambers of YiY_{i} are the chambers of EE that are 𝔞\mathfrak{a}-characteristic for 𝔠i\mathfrak{c}_{i}. The YiY_{i}’s are convex in the sense that if C,CYiC,C^{\prime}\subset Y_{i}, then any minimal gallery from CC to CC^{\prime} is contained in YiY_{i}, and the restriction of ρ\rho to any of the YiY_{i}’s is an isomorphism.

Proof.

For each ii, let EiE_{i} be the apartment asymptotic to 𝔞\mathfrak{a} and 𝔠i\mathfrak{c}_{i}. Then Yi=EEiY_{i}=E\cap E_{i} is a convex subcomplex of EE consisting of the union of the chambers of EE that are 𝔞\mathfrak{a}-characteristic for 𝔠i\mathfrak{c}_{i}. If CC is a chamber of EE, let xy\overrightarrow{xy} be a line segment in ρ(C)\rho(C) in a direction opposite to 𝔞\mathfrak{a}. We can pull it back under ρ\rho to a line segment in CC which points in the direction of a chamber 𝔠iE\mathfrak{c}_{i}\subset\partial_{\infty}E. Then CC is an 𝔞\mathfrak{a}-characteristic chamber for 𝔠i\mathfrak{c}_{i} and CYiC\subset Y_{i}. ∎

Even when CC is not 𝔞\mathfrak{a}-characteristic for 𝔠\mathfrak{c}, the direction DC(𝔠)D_{C}(\mathfrak{c}) still tells us about ρ|x+𝔠\rho|_{x+\mathfrak{c}} for xCx\in C. The following lemma strengthens Lemma 4.5.

Lemma 4.9.

Suppose that 𝔠\mathfrak{c} is a chamber in X\partial_{\infty}X, that CC is a chamber of XX, and x0Cx_{0}\in C. Let CC^{\prime} be a chamber which intersects the sector x0+𝔠x_{0}+\mathfrak{c}. Then either DC(𝔠)=DC(𝔠)D_{C}(\mathfrak{c})=D_{C^{\prime}}(\mathfrak{c}) or dcomb(𝔞,DC(𝔠))<dcomb(𝔞,DC(𝔠))d_{\text{comb}}(\mathfrak{a},D_{C}(\mathfrak{c}))<d_{\text{comb}}(\mathfrak{a},D_{C^{\prime}}(\mathfrak{c})).

Proof.

We proceed similarly to [AB08, 11.63(2)].

Let xCx\in C^{\prime} be a point in Cx0+𝔠C^{\prime}\cap x_{0}+\mathfrak{c}. We may choose xx so that the geodesic segment x0x\overrightarrow{x_{0}x} never crosses two walls simultaneously. Then x0x\overrightarrow{x_{0}x} passes through chambers C=C0,,Cl=CC=C_{0},\dots,C_{l}=C^{\prime} which all meet x0+𝔠x_{0}+\mathfrak{c} and which form a minimal gallery in XX. For each ii, let xix_{i} be a point on x0x\overrightarrow{x_{0}x} which lies on the interior of CiC_{i}.

We proceed inductively. Suppose that the lemma is true for C=C0,,CiC^{\prime}=C_{0},\dots,C_{i} and consider C=Ci+1C^{\prime}=C_{i+1}.

Let EE be an apartment containing CiC_{i} and asymptotic to 𝔞\mathfrak{a}. Let AA be the common panel between CiC_{i} and Ci+1C_{i+1} and let HH be the wall of EE containing AA. Let E+EE^{+}\subset E be the half-apartment bounded by HH which is asymptotic to 𝔞\mathfrak{a} and let EEE^{-}\subset E be the opposite half-apartment.

We consider two cases: CiE+C_{i}\subset E^{+} and CiEC_{i}\subset E^{-}.

If CiE+C_{i}\subset E^{+}, let EE^{\prime} be a ramification of EE (possibly EE itself) which contains E+E^{+} and Ci+1C_{i+1}. This is an apartment asymptotic to 𝔞\mathfrak{a}, so by the definition of ρ\rho, the restriction ρ|E\rho|_{E^{\prime}} is an isomorphism fixing EFE^{\prime}\cap F pointwise. This map sends the line segment xixi+1\overrightarrow{x_{i}x_{i+1}} to the line segment ρ(xi)ρ(xi+1)\overrightarrow{\rho(x_{i})\rho(x_{i+1})}. Since xixi+1\overrightarrow{x_{i}x_{i+1}} is a line segment in the direction of an interior point of 𝔠\mathfrak{c}, this implies that

(1) DCi(𝔠)=DCi+1(𝔠)D_{C_{i}}(\mathfrak{c})=D_{C_{i+1}}(\mathfrak{c})

If CiEC_{i}\subset E^{-}, then we have two possibilities: either Ci+1EC_{i+1}\subset E or Ci+1EC_{i+1}\not\subset E. If Ci+1EC_{i+1}\subset E, then the argument above, applied to EE, shows that DCi(𝔠)=DCi+1(𝔠)D_{C_{i}}(\mathfrak{c})=D_{C_{i+1}}(\mathfrak{c}). Otherwise, let EE^{\prime} be a ramification of EE which contains EE^{-} and Ci+1C_{i+1} and let D=EED=E^{\prime}\smallsetminus E^{-}. Then DE+D\cup E^{+} is an apartment asymptotic to 𝔞\mathfrak{a}, so ρ|D\rho|_{D} is an isomorphism. Likewise, ρ|E\rho|_{E^{-}} is an isomorphism. In fact, the restriction of ρ\rho to E=EDE^{\prime}=E^{-}\cup D is a map EFE^{\prime}\to F which “folds” EE^{\prime} along HH, sending both EE^{-} and DD to ρ(E)\rho(E^{-}).

If s:FFs:F\to F is the reflection fixing ρ(H)\rho(H),

DCi+1(𝔠)=s(DCi(𝔠)).D_{C_{i+1}}(\mathfrak{c})=s_{\infty}(D_{C_{i}}(\mathfrak{c})).

But x0x\overrightarrow{x_{0}x} passes from EE^{-} to E+E^{+}, so 𝔠E+\mathfrak{c}\subset\partial_{\infty}E^{+} and DCi(𝔠)D_{C_{i}}(\mathfrak{c}) is on the same side of ρ(H)\partial_{\infty}\rho(H) as 𝔞\mathfrak{a}. Therefore,

(2) dcomb(𝔞,DCi(𝔠))<dcomb(𝔞,DCi+1(𝔠)).d_{\text{comb}}(\mathfrak{a},D_{C_{i}}(\mathfrak{c}))<d_{\text{comb}}(\mathfrak{a},D_{C_{i+1}}(\mathfrak{c})).

Either (1) or (2) holds for each ii. The lemma follows by induction. ∎

We will also define some families of subsets of XX and XX_{\infty}. Our argument is essentially a quantitative version of Morse theory, so for each point xXx\in X with h(x)0h(x)\geq 0, we will define a set Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) of downward directions, the downward link at infinity and a map from that set to ZZ. By showing that the set of downward directions is highly connected, we will show that ZZ is highly connected.

For any xXx\in X, let S(x)S(x) be the union of the apartments EE such that xEx\in E and 𝔞E\mathfrak{a}\subset\partial_{\infty}E. Let

Lk(x)=S(x)X0(𝔞).\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)=\partial_{\infty}S(x)\cap{X_{\infty}^{0}(\mathfrak{a})}.

The following properties of Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) will be helpful:

Lemma 4.10.
  1. (1)

    If CC is a chamber of XX and xx is in the interior of CC, then 𝔠\mathfrak{c} is a chamber of Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) if and only if CC is 𝔞\mathfrak{a}-characteristic for 𝔠\mathfrak{c}.

  2. (2)

    If CC is a chamber of XX, xx is in the interior of CC, and 𝔠,𝔠Lk(x)\mathfrak{c},\mathfrak{c}^{\prime}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x), then 𝔠\mathfrak{c} and 𝔠\mathfrak{c}^{\prime} point in the same direction at CC.

  3. (3)

    If xx+𝔞x^{\prime}\in x+\mathfrak{a}, then Lk(x)Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime}).

  4. (4)

    If QXQ\subset X is a bounded subset, then there is an xXx\in X such that d(Q,x)diamQd(Q,x)\lesssim\operatorname{diam}Q and xq+𝔞x\in q+\mathfrak{a} for any qQq\in Q.

  5. (5)

    If r:[0,)r:[0,\infty) is a unit-speed ray emanating from xx in the direction of a point σLk(x)\sigma\in\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x), then

    h(r(t))=h(x)+tcosd(τ,σ).h(r(t))=h(x)+t\cos d(\tau,\sigma).

    Furthermore, there is an ϵ>0\epsilon>0 depending on XX and p(τ)p(\tau) such that cosd(τ,σ)>ϵ-\cos d(\tau,\sigma)>\epsilon.

Proof.

The first property follows from the definition of Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) and the fact that CC is an 𝔞\mathfrak{a}-characteristic chamber for 𝔠\mathfrak{c} if and only if 𝔞\mathfrak{a} and 𝔠\mathfrak{c} are opposite and the unique apartment asymptotic to 𝔞\mathfrak{a} and 𝔠\mathfrak{c} contains CC.

If xx is in the interior of CC and 𝔠,𝔠Lk(x)\mathfrak{c},\mathfrak{c}^{\prime}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x), then CC is 𝔞\mathfrak{a}-characteristic for 𝔠\mathfrak{c} and 𝔠\mathfrak{c}^{\prime}. Consequently, DC(𝔠)D_{C}(\mathfrak{c}) and DC(𝔠)D_{C}(\mathfrak{c}^{\prime}) are both the chamber of F\partial_{\infty}F opposite to 𝔞\mathfrak{a}, so 𝔠\mathfrak{c} and 𝔠\mathfrak{c}^{\prime} point in the same direction at CC.

For the third property, we show that S(x)S(x)S(x)\subset S(x^{\prime}). If yS(x)y\in S(x), then there is an apartment containing xx and yy and asymptotic to 𝔞\mathfrak{a}. Since xx+𝔞x^{\prime}\in x+\mathfrak{a}, xx^{\prime} lies in this apartment as well. It follows that Lk(x)Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime}).

To prove the fourth property, for all qQq\in Q, let rq:[0,)Xr_{q}:[0,\infty)\to X be a ray emanating from qq in the direction of the barycenter of 𝔞\mathfrak{a}. Let EE be an apartment asymptotic to 𝔞\mathfrak{a} that intersects QQ nontrivially. Then d(q,E)diamQd(q,E)\leq\operatorname{diam}Q for any qQq\in Q, so by Lemma 4.6.3 of [KL97], there is a cc such that if tcdiamQt\geq c\operatorname{diam}Q, then rq(t)Er_{q}(t)\in E. In particular, V=qrq(t)+𝔞V=\bigcap_{q}r_{q}(t)+\mathfrak{a} is a sector in EE that satisfies Vq+𝔞V\subset q+\mathfrak{a} for all qq and d(V,Q)diamQd(V,Q)\lesssim\operatorname{diam}Q. Choose xVx\in V.

Finally, if rr is a ray in the direction of σ\sigma, let EE be an apartment which contains xx and is asymptotic to 𝔞\mathfrak{a} and to σ\sigma. Then rr is a geodesic ray in EE, which makes an angle of d(τ,σ)d(\tau,\sigma) with the ray emanating from xx in the direction of τ\tau. The formula for h(r(t))h(r(t)) follows by trigonometry.

To bound d(τ,σ)d(\tau,\sigma), consider

m=maxθ𝔞d(τ,θ).m=\max_{\theta\in\mathfrak{a}}d(\tau,\theta).

If σ¯\bar{\sigma} is the direction opposite to σ\sigma in E\partial_{\infty}E, then by the definition of Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x), we have σ¯𝔞\bar{\sigma}\in\mathfrak{a}, so d(τ,σ)=πd(τ,σ¯)πmd(\tau,\sigma)=\pi-d(\tau,\bar{\sigma})\geq\pi-m. We claim that m<π/2m<\pi/2.

By Lemma 4.1 of [BW11], the diameter of 𝔞\mathfrak{a} is at most π/2\pi/2, and if the diameter is equal to π/2\pi/2, then 𝔞\mathfrak{a} is a nontrivial spherical join and XX is a nontrivial product of buildings. Furthermore, if θ𝔞\theta\in\mathfrak{a} is such that d(τ,θ)=π/2d(\tau,\theta)=\pi/2, then we can write X=X1×X2X=X_{1}\times X_{2} such that τ(X1),θ(X2)\tau\in(X_{1})_{\infty},\theta\in(X_{2})_{\infty}. This contradicts the hypothesis that τ\tau is not parallel to a factor of XX, so m<π/2m<\pi/2 and cosd(τ,σ)cosm>0-\cos d(\tau,\sigma)\geq-\cos m>0. ∎

4.3. Apartments in X0(𝔞){X_{\infty}^{0}(\mathfrak{a})}

In this section, we use the tools of the previous section to construct apartments in X0(𝔞){X_{\infty}^{0}(\mathfrak{a})}; in the next section, we will use these apartments to contract spheres in X0(𝔞){X_{\infty}^{0}(\mathfrak{a})}. First, we show that every chamber in X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} is part of some apartment in X0(𝔞){X_{\infty}^{0}(\mathfrak{a})}:

Lemma 4.11.

Suppose that 𝔠\mathfrak{c} is a chamber of XX_{\infty} opposite to 𝔞\mathfrak{a} and suppose that CC is an 𝔞\mathfrak{a}-characteristic chamber for 𝔠\mathfrak{c}. There is an apartment EE containing CC such that EE is asymptotic to 𝔠\mathfrak{c} and every chamber of E\partial_{\infty}E is opposite to 𝔞\mathfrak{a}.

Furthermore, there is a c>0c>0 depending only on XX and an 𝔞\mathfrak{a}-characteristic chamber C𝔟EC_{\mathfrak{b}}\subset E for each chamber 𝔟E\mathfrak{b}\subset\partial_{\infty}E such that C𝔠=CC_{\mathfrak{c}}=C and

diam𝔟EC𝔟c.\operatorname{diam}\bigcup_{\mathfrak{b}\subset\partial_{\infty}E}C_{\mathfrak{b}}\leq c.

We will prove this lemma by starting with an apartment EXE\subset X, then producing a series of ramifications of EE so that more and more chambers of E\partial_{\infty}E are opposite to 𝔞\mathfrak{a}. Since XX is thick, if 𝔠\mathfrak{c} is a chamber of E\partial_{\infty}E which is not opposite to 𝔞\mathfrak{a}, then there is some ramification EE^{\prime} of EE that replaces 𝔠\mathfrak{c} with a chamber that is farther (in XX_{\infty}) from 𝔞\mathfrak{a}. This might replace a chamber of E\partial_{\infty}E which is already opposite to 𝔞\mathfrak{a} with a chamber which is not, but we avoid this by ensuring that EE^{\prime} contains the same 𝔞\mathfrak{a}-characteristic chambers as EE.

The following lemma produces these ramifications:

Lemma 4.12.

Let EE be an apartment of XX and let 𝔠=𝔠1,,𝔠k\mathfrak{c}=\mathfrak{c}_{1},\dots,\mathfrak{c}_{k} be chambers of E\partial_{\infty}E which are opposite to 𝔞\mathfrak{a}. Let CiEC_{i}\subset E be a 𝔞\mathfrak{a}-characteristic chamber for 𝔠i\mathfrak{c}_{i} for each ii. Let 𝔟\mathfrak{b} be a chamber of E\partial_{\infty}E, distinct from the 𝔠i\mathfrak{c}_{i}’s, which is adjacent to 𝔠\mathfrak{c}. There is a ramification E0E_{0} of EE such that if ϕ:EE0\phi:E\to E_{0} is the isomorphism fixing EE0E\cap E_{0} pointwise, then

  • CiEE0C_{i}\subset E\cap E_{0} for all ii (and thus ϕ(𝔠i)\phi_{\infty}(\mathfrak{c}_{i}) is opposite to 𝔞\mathfrak{a}),

  • ϕ(𝔟)\phi_{\infty}(\mathfrak{b}) is opposite to 𝔞\mathfrak{a}, and

  • there is an 𝔞\mathfrak{a}-characteristic chamber B0E0B_{0}\subset E_{0} for ϕ(𝔟)\phi_{\infty}(\mathfrak{b}) such that d(B0,Ci)diamCid(B_{0},\bigcup C_{i})\lesssim\operatorname{diam}\bigcup C_{i}.

Proof.

Let C=C1C=C_{1} and let x0Cx_{0}\in C. Let HH be a wall in EE such that H\partial_{\infty}H separates 𝔟\mathfrak{b} and 𝔠\mathfrak{c}. Let M,MEM,M^{\prime}\subset E be the half-apartments of EE bounded by HH. By translating HH and possibly switching MM and MM^{\prime}, we may arrange that

  • 𝔠M\mathfrak{c}\in\partial_{\infty}M and 𝔟M,\mathfrak{b}\in\partial_{\infty}M^{\prime},

  • CiMC_{i}\subset M for all ii, and

  • d(H,C)diam(Ci).d(H,C)\lesssim\operatorname{diam}(\bigcup C_{i}).

We claim that there is a ramification E0E_{0} of EE which contains MM and satisfies the conditions of the lemma.

By our choice of HH, the intersection x0+𝔟Mx_{0}+\mathfrak{b}\cap M^{\prime} is a sector of EE, and we can choose Bx0+𝔟MB\subset x_{0}+\mathfrak{b}\cap M^{\prime} to be a chamber which borders HH and satisfies d(x0,B)diam(Ci).d(x_{0},B)\lesssim\operatorname{diam}(\bigcup C_{i}). Let AA be the panel of HH bordering BB, let DMD\subset M be the chamber of EE adjacent to BB along AA, and let BB^{\prime} be a chamber adjacent to AA and distinct from BB and DD. Let EE^{\prime} be a ramification of EE that contains BB^{\prime} and let ϕ:EE\phi:E\to E^{\prime} be the isomorphism fixing EEE\cap E^{\prime}. We claim that either the lemma is satisfied for E0=EE_{0}=E and B0=BB_{0}=B or it is satisfied for E0=EE_{0}=E^{\prime} and B0=BB_{0}=B^{\prime}.

Since 𝔞\mathfrak{a} is opposite to DC(𝔠)D_{C}(\mathfrak{c}) and DC(𝔟)D_{C}(\mathfrak{b}) is adjacent to DC(𝔠)D_{C}(\mathfrak{c}),

dcomb(𝔞,DC(𝔟))=dcomb(𝔞,DC(𝔠))1.d_{\text{comb}}(\mathfrak{a},D_{C}(\mathfrak{b}))=d_{\text{comb}}(\mathfrak{a},D_{C}(\mathfrak{c}))-1.

Lemma 4.9 implies that either DB(𝔟)D_{B}(\mathfrak{b}) is opposite to 𝔞\mathfrak{a} or DB(𝔟)=DC(𝔟)D_{B}(\mathfrak{b})=D_{C}(\mathfrak{b}). By Lemma 4.4, DBD_{B} and DCD_{C} are type-preserving isomorphisms from E\partial_{\infty}E to F\partial_{\infty}F, so if DB(𝔟)=DC(𝔟)D_{B}(\mathfrak{b})=D_{C}(\mathfrak{b}), then DB=DCD_{B}=D_{C}, and BB is 𝔞\mathfrak{a}-characteristic for 𝔠\mathfrak{c}. So BB is 𝔞\mathfrak{a}-characteristic for either 𝔟\mathfrak{b} or 𝔠\mathfrak{c}. In the first case, the lemma is satisfied for E0=EE_{0}=E and B0=BB_{0}=B.

Likewise, if 𝔟=ϕ(𝔟)\mathfrak{b}^{\prime}=\phi_{\infty}(\mathfrak{b}), then DC(𝔟)=DC(𝔟)D_{C}(\mathfrak{b}^{\prime})=D_{C}(\mathfrak{b}) is adjacent to DC(𝔠)D_{C}(\mathfrak{c}) and Bx0+𝔟B^{\prime}\subset x_{0}+\mathfrak{b}^{\prime}, so BB^{\prime} is 𝔞\mathfrak{a}-characteristic for either 𝔟\mathfrak{b}^{\prime} or 𝔠\mathfrak{c}. In the first case, the lemma is satisfied for E0=EE_{0}=E^{\prime} and B0=BB_{0}=B^{\prime}.

Suppose by way of contradiction that BB and BB^{\prime} are both 𝔞\mathfrak{a}-characteristic for 𝔠\mathfrak{c}. The union of the set of chambers of XX that are 𝔞\mathfrak{a}-characteristic for 𝔠\mathfrak{c} is the unique apartment E𝔞,𝔠E_{\mathfrak{a},\mathfrak{c}} asymptotic to 𝔞\mathfrak{a} and 𝔠\mathfrak{c}, so in particular, it is convex. It contains BB and CC, so it contains DD as well. But then BB, BB^{\prime}, and DD are distinct chambers of E𝔞,𝔠E_{\mathfrak{a},\mathfrak{c}} which are all adjacent to the same panel. This is impossible. ∎

Proof of Lemma 4.11.

Let E𝔞,𝔠XE_{\mathfrak{a},\mathfrak{c}}\subset X be the apartment spanned by 𝔞\mathfrak{a} and 𝔠\mathfrak{c}, so that CE𝔞,𝔠C\subset E_{\mathfrak{a},\mathfrak{c}}. By applying Lemma 4.12 to E𝔞,𝔠E_{\mathfrak{a},\mathfrak{c}} repeatedly, we can construct an apartment EE such that for any chamber 𝔟E\mathfrak{b}\in\partial_{\infty}E, there is an 𝔞\mathfrak{a}-characteristic chamber C𝔟C_{\mathfrak{b}} for 𝔟\mathfrak{b}, and

diam𝔟EC𝔟\operatorname{diam}\bigcup_{\mathfrak{b}\subset\partial_{\infty}E^{\prime}}C_{\mathfrak{b}}

is bounded. ∎

In fact, we can find many apartments in X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} simultaneously:

Lemma 4.13.

Suppose that EE is an apartment of XX and suppose that for each chamber 𝔠E\mathfrak{c}\subset\partial_{\infty}E there is a chamber C𝔠EC_{\mathfrak{c}}\subset E which is 𝔞\mathfrak{a}-characteristic for 𝔠\mathfrak{c} and a point x𝔠C𝔠x_{\mathfrak{c}}\in C_{\mathfrak{c}}. Let 𝔟\mathfrak{b} and ¯𝔟\bar{}\mathfrak{b} be two opposite chambers in E\partial_{\infty}E. Suppose that CC is a chamber of XX and xx is a point in the interior of CC such that xx𝔟+𝔟x\in x_{\mathfrak{b}}+\mathfrak{b} and C𝔠x+¯𝔟C_{\mathfrak{c}}\subset x+\bar{}\mathfrak{b} for all 𝔠E\mathfrak{c}\subset\partial_{\infty}E. Then there is an xx+𝔞x^{\prime}\in x+\mathfrak{a} such that

d(x,x)diam𝔠EC𝔠d(x,x^{\prime})\lesssim\operatorname{diam}\bigcup_{\mathfrak{c}\subset\partial_{\infty}E}C_{\mathfrak{c}}

and for every chamber 𝔡Lk(x)\mathfrak{d}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x),

  • 𝔡\mathfrak{d} is opposite to ¯𝔟\bar{}\mathfrak{b},

  • if E𝔡,¯𝔟E_{\mathfrak{d},\bar{}\mathfrak{b}} is the apartment spanned by 𝔡\mathfrak{d} and ¯𝔟\bar{}\mathfrak{b}, then E𝔡,¯𝔟Lk(x)\partial_{\infty}E_{\mathfrak{d},\bar{}\mathfrak{b}}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime}).

Proof.

Suppose that 𝔡Lk(x)\mathfrak{d}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x). Then CC is 𝔞\mathfrak{a}-characteristic for 𝔟\mathfrak{b} and 𝔡\mathfrak{d}, so 𝔟\mathfrak{b} and 𝔡\mathfrak{d} point in the same direction at CC. Since 𝔟\mathfrak{b} and ¯𝔟\bar{}\mathfrak{b} point in opposite directions at CC, we conclude that CC is ¯𝔟\bar{}\mathfrak{b}-characteristic for 𝔡\mathfrak{d}. Thus, ¯𝔟\bar{}\mathfrak{b} and 𝔡\mathfrak{d} are opposite and CE𝔡,¯𝔟C\subset E_{\mathfrak{d},\bar{}\mathfrak{b}}.

In particular, x+¯𝔟E𝔡,¯𝔟x+\bar{}\mathfrak{b}\subset E_{\mathfrak{d},\bar{}\mathfrak{b}}, so C𝔠E𝔡,¯𝔟C_{\mathfrak{c}}\subset E_{\mathfrak{d},\bar{}\mathfrak{b}} for all 𝔠E\mathfrak{c}\subset\partial_{\infty}E. Let ϕ:E𝔡,¯𝔟E\phi:E_{\mathfrak{d},\bar{}\mathfrak{b}}\to E be the isomorphism fixing E𝔡,¯𝔟EE_{\mathfrak{d},\bar{}\mathfrak{b}}\cap E pointwise and suppose that 𝔠E𝔡,¯𝔟\mathfrak{c}^{\prime}\subset\partial_{\infty}E_{\mathfrak{d},\bar{}\mathfrak{b}}. If 𝔠=ϕ(𝔠)\mathfrak{c}=\phi_{\infty}(\mathfrak{c}^{\prime}), then C𝔠C_{\mathfrak{c}} is an 𝔞\mathfrak{a}-characteristic chamber for 𝔠\mathfrak{c}^{\prime}, so 𝔠Lk(x𝔠)\mathfrak{c}^{\prime}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x_{\mathfrak{c}}).

By Lemma 4.10(3) and (4), there is an

x𝔠Ex𝔠+𝔞.x^{\prime}\in\bigcap_{\mathfrak{c}\subset\partial_{\infty}E}x_{\mathfrak{c}}+\mathfrak{a}.

such that xx+𝔞x^{\prime}\in x+\mathfrak{a} and Lk(x𝔠)Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x_{\mathfrak{c}})\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime}) for every 𝔠E\mathfrak{c}\subset\partial_{\infty}E. ∎

Combining Lemmas 4.13 and 4.11 we get:

Lemma 4.14.

For any xXx\in X, there is a chamber 𝔡X\mathfrak{d}\subset X_{\infty} opposite to 𝔞\mathfrak{a} and an xx+𝔞x^{\prime}\in x+\mathfrak{a} such that

  • if 𝔠Lk(x)\mathfrak{c}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) then 𝔡\mathfrak{d} is opposite to 𝔠\mathfrak{c},

  • if 𝔠Lk(x)\mathfrak{c}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) and E𝔠,𝔡E_{\mathfrak{c},\mathfrak{d}} is the apartment spanned by 𝔠\mathfrak{c} and 𝔡\mathfrak{d}, then E𝔠,𝔡Lk(x)\partial_{\infty}E_{\mathfrak{c},\mathfrak{d}}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime}), and

  • d(x,x)1d(x,x^{\prime})\lesssim 1.

Proof.

Let 𝔟Lk(x)\mathfrak{b}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) and let EE be the unique apartment asymptotic to 𝔞\mathfrak{a} and 𝔟\mathfrak{b}. Since 𝔟Lk(x)\mathfrak{b}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x), we have xEx\in E. We may perturb xx in the direction of 𝔞\mathfrak{a} to ensure that xx is in the interior of some chamber CC of EE; this doesn’t change Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x). Let rr be a unit-speed ray emanating from xx in the direction of the barycenter of 𝔞\mathfrak{a} and let 0<θ<π/20<\theta<\pi/2 be the minimum angle between the barycenter of 𝔞\mathfrak{a} and any point on its boundary. Let cc be the constant in Lemma 4.11 and let t>csinθt>\frac{c}{\sin{\theta}}, so that

BE(r(t),c)x+𝔞,B_{E}(r(t),c)\subset x+\mathfrak{a},

where BE(r(t),c)B_{E}(r(t),c) is the ball in EE with center r(t)r(t) and radius cc. Let x0=r(t)x_{0}=r(t).

Let C0EC_{0}\subset E be a chamber such that x0C0x_{0}\in C_{0}. Since C0EC_{0}\subset E, it is 𝔞\mathfrak{a}-characteristic for 𝔟\mathfrak{b}. By Lemma 4.11, there is an apartment EE^{\prime} and a collection of 𝔞\mathfrak{a}-characteristic chambers C𝔠EC_{\mathfrak{c}}\subset E^{\prime} for 𝔠E\mathfrak{c}\subset\partial_{\infty}E^{\prime} such that x0+𝔟Ex_{0}+\mathfrak{b}\subset E^{\prime} and

𝔠EC𝔠BE(x0,c).\bigcup_{\mathfrak{c}\subset\partial_{\infty}E^{\prime}}C_{\mathfrak{c}}\subset B_{E^{\prime}}(x_{0},c).

Let ¯𝔟\bar{}\mathfrak{b} be the chamber of E\partial_{\infty}E^{\prime} opposite to 𝔟\mathfrak{b}. We claim that x+¯𝔟x+\bar{}\mathfrak{b} contains all of the C𝔠C_{\mathfrak{c}}’s.

Let ϕ:EE\phi:E\to E^{\prime} be the isomorphism fixing EEE\cap E^{\prime} pointwise. Then ϕ\phi fixes CC and C0C_{0} and sends 𝔞\mathfrak{a} to ¯𝔟\bar{}\mathfrak{b}, so ϕ(x+𝔞)=x+¯𝔟\phi(x+\mathfrak{a})=x+\bar{}\mathfrak{b} and ϕ(BE(x0,c))=BE(x0,c)\phi(B_{E}(x_{0},c))=B_{E^{\prime}}(x_{0},c). Therefore,

𝔠EC𝔠BE(x0,t)x+¯𝔟.\bigcup_{\mathfrak{c}\subset\partial_{\infty}E^{\prime}}C_{\mathfrak{c}}\subset B_{E^{\prime}}(x_{0},t)\subset x+\bar{}\mathfrak{b}.

By applying Lemma 4.13 to EE^{\prime}, we obtain an xx^{\prime} that satisfies the required properties and has

d(x,x)diam𝔠EC𝔠1.d(x,x^{\prime})\lesssim\operatorname{diam}\bigcup_{\mathfrak{c}\subset\partial_{\infty}E^{\prime}}C_{\mathfrak{c}}\lesssim 1.

We can also use these techniques to construct (n1)(n-1)-spheres in ZZ which are homotopically nontrivial in ZZ. This generalizes results of Bux and Wortman [BW07] on buildings acted on by SS-arithmetic groups to arbitrary euclidean buildings.

Lemma 4.15.

For any r>0r>0, there is a map α:Sn1Z\alpha:S^{n-1}\to Z such that α\alpha is homotopically nontrivial in Nr(Z)N_{r}(Z), where Nr(Z)N_{r}(Z) is the rr-neighborhood of ZZ.

Proof.

Let CC be a chamber of XX such that minxCh(x)>r\min_{x\in C}h(x)>r. Let EE be an apartment containing CC and asymptotic to 𝔞\mathfrak{a}. If 𝔠E\mathfrak{c}\subset\partial_{\infty}E is the chamber of E\partial_{\infty}E opposite to 𝔞\mathfrak{a}, then CC is 𝔞\mathfrak{a}-characteristic for 𝔠\mathfrak{c}. Using Lemma 4.12, we can construct an apartment EE^{\prime} such that CEC\subset E^{\prime} and EX0(𝔞)\partial_{\infty}E^{\prime}\subset{X_{\infty}^{0}(\mathfrak{a})}. In particular, the set of points B={xEh(x)0}B=\{x\in E^{\prime}\mid h(x)\geq 0\} is convex and compact and contains CC, so ZEZ\cap E^{\prime} is bilipschitz equivalent to the (n1)(n-1)-sphere. Let α:Sn1ZE\alpha:S^{n-1}\to Z\cap E^{\prime} be a Lipschitz homeomorphism. We claim that α\alpha is homotopically nontrivial in Nr(Z)N_{r}(Z).

Let β:DnE\beta:D^{n}\to E be a homeomorphism from DnBD^{n}\to B which extends α\alpha. This has degree 1 on any point in the interior of CC. By way of contradiction, suppose that β:DnNr(Z)\beta^{\prime}:D^{n}\to N_{r}(Z) is another extension of α\alpha. Then we can glue β\beta and β\beta^{\prime} together to get a map γ:SnX\gamma:S^{n}\to X. Since β\beta^{\prime} avoids CC, this map has degree 1 on any point in the interior of CC. Since XX is CAT(0), however, it is contractible, so γ\gamma must be null-homotopic, and γ\gamma sends the fundamental class of SnS^{n} to an nn-boundary in XX. This contradicts the fact that this map has degree 1 on any point in the interior of CC, because XX is nn-dimensional, and any nn-boundary must be trivial. ∎

4.4. (n2)(n-2)-connectivity for X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} and constructing Ω\Omega_{\infty}

The lemmas of the previous section will let us prove that X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} is (n2)(n-2)-connected and construct a Lipschitz map

Ω:ΔZ(n1)X0(𝔞)\Omega_{\infty}:\Delta_{Z}^{(n-1)}\to X_{\infty}^{0}(\mathfrak{a})

which we will use to construct Ω\Omega.

Let ΔZ\Delta_{Z} be the infinite-dimensional simplex with vertex set ZZ. As before, we denote the simplex of ΔZ\Delta_{Z} with vertices z0,,zkz_{0},\dots,z_{k} by z0,,zk\langle z_{0},\dots,z_{k}\rangle. If Δ\Delta is a simplex of ΔZ\Delta_{Z}, we let 𝒱(Δ)Z\mathcal{V}(\Delta)\subset Z be the vertex set of Δ\Delta.

The main lemma of this section is the following:

Lemma 4.16.

There is a cellular map

Ω:ΔZ(n1)X0(𝔞),\Omega_{\infty}:\Delta_{Z}^{(n-1)}\to X_{\infty}^{0}(\mathfrak{a}),

a c>0c>0 depending on XX, and a family of points xΔXx_{\Delta}\in X, one for each simplex ΔΔZ(n1)\Delta\subset\Delta_{Z}^{(n-1)}, such that

  1. (1)

    mass(Ω(Δ))c\operatorname{mass}(\Omega_{\infty}(\Delta))\leq c

  2. (2)

    h(xΔ)0h(x_{\Delta})\geq 0,

  3. (3)

    Ω(Δ)Lk(xΔ),\Omega_{\infty}(\Delta)\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x_{\Delta}),

  4. (4)

    if ΔΔ\Delta^{\prime}\subset\Delta, then xΔxΔ+𝔞x_{\Delta}\in x_{\Delta^{\prime}}+\mathfrak{a},

  5. (5)

    and d(xΔ,𝒱(Δ))diam𝒱(Δ)+1d(x_{\Delta},\mathcal{V}(\Delta))\lesssim\operatorname{diam}\mathcal{V}(\Delta)+1 (consequently, h(xΔ)diam𝒱(Δ)+1h(x_{\Delta})\lesssim\operatorname{diam}\mathcal{V}(\Delta)+1),

Furthermore, for any zZz\in Z, we have xz=zx_{\langle z\rangle}=z.

The first condition is essentially a bound on the filling functions of X0(𝔞){X_{\infty}^{0}(\mathfrak{a})}. The next three conditions ensure that the map ixΔi_{x_{\Delta}} (as defined in the proof sketch at the beginning of the section) is defined on Ω(Δ)\Omega_{\infty}(\Delta) and that its Lipschitz constant is diam𝒱(Δ)\lesssim\operatorname{diam}\mathcal{V}(\Delta). In order to construct Ω\Omega in the next section, we will glue maps of the form ixΔΩ|Δi_{x_{\Delta}}\circ\Omega_{\infty}|_{\Delta}, and we will use the last condition to perform this gluing.

First, we prove that X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} is (n2)(n-2)-connected.

Lemma 4.17.

If k<n1k<n-1, there is a c>0c>0 such that for every xXx\in X, there is a xXx^{\prime}\in X such that xx+𝔞x^{\prime}\in x+\mathfrak{a}, d(x,x)cd(x,x^{\prime})\leq c, and if

α:SkLk(x)(k),\alpha:S^{k}\to\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)^{(k)},

then there is an extension

β:Bk+1Lk(x)(k+1)\beta:B^{k+1}\to\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime})^{(k+1)}

such that LipβcLipα+c\operatorname{Lip}\beta\leq c^{\prime}\operatorname{Lip}\alpha+c^{\prime}.

Consequently, X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} is (n2)(n-2)-connected.

Proof.

Let xXx^{\prime}\in X and 𝔡Lk(x)\mathfrak{d}\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime}) be opposite to every chamber of Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) as in Lemma 4.14. Let uu be the barycenter of 𝔡\mathfrak{d}. There is an ϵ>0\epsilon>0 such that dX(u,v)<πϵd_{X\infty}(u,v)<\pi-\epsilon for any vLk(x)(k)v\in\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)^{(k)}. By our choice of 𝔡\mathfrak{d}, the geodesic from vv to uu is contained in Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime}).

Let

γ:Lk(x)(k)×[0,1]Lk(x)\gamma:\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)^{(k)}\times[0,1]\to\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime})

be the map which sends v×[0,1]v\times[0,1] to the geodesic between vv and uu. This is Lipschitz, with Lipschitz constant depending on ϵ\epsilon. Define β0:Sk×[0,1]Lk(x)\beta_{0}:S^{k}\times[0,1]\to\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime}) by β0(v,t)=γ(α(v),t)\beta_{0}(v,t)=\gamma(\alpha(v),t). This is a null-homotopy of α\alpha, and

Lipβ0(Lipγ)(1+Lipα).\operatorname{Lip}\beta_{0}\leq(\operatorname{Lip}\gamma)(1+\operatorname{Lip}\alpha).

We obtain β\beta by approximating β0\beta_{0} in Lk(x)(k+1)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)^{(k+1)}; this increases the Lipschitz constant by at most a multiplicative factor.

To conclude that X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} is (n2)(n-2)-connected, consider a map α:SkX0(𝔞)\alpha:S^{k}\to{X_{\infty}^{0}(\mathfrak{a})}. This can be approximated by a simplicial map α:SkX0(𝔞)(k)\alpha^{\prime}:S^{k}\to{X_{\infty}^{0}(\mathfrak{a})}^{(k)}. The image of α\alpha^{\prime} has finitely many simplices, and since every simplex of X0(𝔞){X_{\infty}^{0}(\mathfrak{a})} is contained in Lk(y)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(y) for some yy, there is an xXx\in X such that the image of α\alpha^{\prime} is contained in Lk(x)(k)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)^{(k)}. Therefore, α\alpha^{\prime} is null-homotopic in Lk(x)(k)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime})^{(k)} for some xx^{\prime}, and Lk(x)(k)X0(𝔞)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime})^{(k)}\subset{X_{\infty}^{0}(\mathfrak{a})}. ∎

Next, we use this lemma to construct Ω\Omega_{\infty}:

Proof of Lemma 4.16.

We construct Ω\Omega_{\infty} inductively. First, for each zZz\in Z, we let Ω(z)\Omega_{\infty}(z) be an arbitrary vertex of Lk(z)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(z). Then choosing xz=zx_{\langle z\rangle}=z satisfies the conditions of the lemma.

Now suppose that Δ\Delta is a simplex of ΔZ\Delta_{Z} with 1dimΔ=kn11\leq\dim\Delta=k\leq n-1 and suppose that Ω\Omega_{\infty} is defined on Δ\partial\Delta. Then, if

L=ΔΔLk(xΔ),L=\bigcup_{\Delta^{\prime}\subset\Delta}\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x_{\Delta^{\prime}}),

then Ω|Δ\Omega_{\infty}|_{\partial\Delta} is a map with image in LL. By induction, we know that

d(xΔ,𝒱(Δ))kdiam𝒱(Δ)+1,d(x_{\Delta^{\prime}},\mathcal{V}(\Delta^{\prime}))\lesssim_{k}\operatorname{diam}\mathcal{V}(\Delta^{\prime})+1,

with implicit constant depending on kk, so

diam{xΔ}ΔΔkdiam𝒱(Δ)+1.\operatorname{diam}\{x_{\Delta^{\prime}}\}_{\Delta^{\prime}\subset\Delta}\lesssim_{k}\operatorname{diam}\mathcal{V}(\Delta)+1.

By Lemma 4.10(4), there is an x0Xx_{0}\in X such that d(x0,𝒱(Δ))kdiam𝒱(Δ)+1d(x_{0},\mathcal{V}(\Delta))\lesssim_{k}\operatorname{diam}\mathcal{V}(\Delta)+1 and x0xΔ+𝔞x_{0}\in x_{\Delta^{\prime}}+\mathfrak{a} for any face Δ\Delta^{\prime} of Δ\Delta. By Lemma 4.10(3), LLk(x0)L\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x_{0}).

By Lemma 4.17, there is an xXx^{\prime}\in X such that xx0+𝔞x^{\prime}\in x_{0}+\mathfrak{a} and an extension

β:ΔLk(x)(k+1)\beta:\Delta\to\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime})^{(k+1)}

of α\alpha such that Lipβ\operatorname{Lip}\beta and d(x0,x)d(x_{0},x^{\prime}) are bounded by a constant depending on kk. If we define Ω|Δ=β\Omega_{\infty}|_{\Delta}=\beta and xΔ=xx_{\Delta}=x^{\prime}, then

d(xΔ,𝒱(Δ))kdiam𝒱(Δ)+1.d(x_{\Delta},\mathcal{V}(\Delta))\lesssim_{k}\operatorname{diam}\mathcal{V}(\Delta)+1.

Since Δ\Delta is finite-dimensional, we may drop the dependence on kk, and the lemma holds. ∎

4.5. Constructing Ω\Omega

Finally, we construct a map Ω:ΔZ(n1)Z\Omega:\Delta_{Z}^{(n-1)}\to Z satisfying the hypotheses of Lemma 3.2. We will use a family of maps ix:Lk(x)Zi_{x}:\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)\to Z for xX,h(x)0x\in X,h(x)\geq 0.

For any xXx\in X and σX\sigma\in X_{\infty}, there is a unit-speed ray rσ:[0,)Xr_{\sigma}:[0,\infty)\to X emanating from xx and traveling in the direction of σ\sigma. Define

X=X×[0,)/X×{0}X_{\infty}^{*}=X_{\infty}\times[0,\infty)/X_{\infty}\times\{0\}

to be a space of “vectors” based at xx. We can define an exponential map ex:XXe_{x}:X_{\infty}^{*}\to X by letting

ex(σ,t)=rσ(t).e_{x}(\sigma,t)=r_{\sigma}(t).

For each chamber 𝔞\mathfrak{a} of XX_{\infty}, this map sends the open cone 𝔞×[0,)/𝔞×{0}\mathfrak{a}\times[0,\infty)/\mathfrak{a}\times\{0\} to a sector corresponding to 𝔞\mathfrak{a}; we give XX_{\infty}^{*} a metric so that this is an isometry. This makes exe_{x} a distance-decreasing map. Note also that, by the convexity of the distance function on XX, we have

d(ex(σ,t),ex(σ,t))d(x,x).d(e_{x}(\sigma,t),e_{x^{\prime}}(\sigma,t))\leq d(x,x^{\prime}).

We can use exe_{x} to construct a map from Lk(x)\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) to ZZ:

Lemma 4.18.

Let xXx\in X be such that h(x)0h(x)\geq 0. Then there is a map ix:Lk(x)Zi_{x}:\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)\to Z given by

ix(σ)=ex(σ,h(x)cosd(τ,σ)).i_{x}(\sigma)=e_{x}(\sigma,\frac{-h(x)}{\cos d(\tau,\sigma)}).

This map has Lipschitz constant h(x)\lesssim h(x), with implicit constant depending on XX and p(τ)p(\tau).

Proof.

By Lemma 4.10(5),

h(ex(σ,t))=h(x)+tcosd(τ,σ)h(e_{x}(\sigma,t))=h(x)+t\cos d(\tau,\sigma)

for any σLk(x)\sigma\in\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x) and there is an ϵ\epsilon such that cosd(τ,σ)ϵ>0-\cos d(\tau,\sigma)\geq\epsilon>0. The lemma follows. ∎

Furthermore, the map (x,σ)ix(σ)(x,\sigma)\mapsto i_{x}(\sigma) is locally Lipschitz:

Lemma 4.19.

Let x,xXx,x^{\prime}\in X be such that h(x),h(x)0h(x),h(x^{\prime})\geq 0. Let σ,σLk(x)Lk(x)\sigma,\sigma^{\prime}\in\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x)\cap\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x^{\prime}). Then there is a c>0c>0 depending on XX such that

d(ix(σ),ix(σ))cd(x,x)+ch(x)d(σ,σ).d(i_{x}(\sigma),i_{x^{\prime}}(\sigma^{\prime}))\leq cd(x,x^{\prime})+ch(x)d(\sigma,\sigma^{\prime}).
Proof.

By the previous lemma and the remark before it, there is a c0>0c_{0}>0 such that

d(ix(σ),ix(σ))\displaystyle d(i_{x}(\sigma),i_{x^{\prime}}(\sigma^{\prime})) d(ix(σ),ix(σ))+d(ix(σ),ix(σ))\displaystyle\leq d(i_{x}(\sigma),i_{x}(\sigma^{\prime}))+d(i_{x}(\sigma^{\prime}),i_{x^{\prime}}(\sigma^{\prime}))
c0h(x)d(σ,σ)+d(ex(σ,h(x)cosd(τ,σ)),ex(σ,h(x)cosd(τ,σ)))\displaystyle\leq c_{0}h(x)d(\sigma,\sigma^{\prime})+d\Bigl{(}e_{x}\bigl{(}\sigma^{\prime},\frac{-h(x)}{\cos d(\tau,\sigma^{\prime})}\bigr{)},e_{x^{\prime}}\bigl{(}\sigma^{\prime},\frac{-h(x^{\prime})}{\cos d(\tau,\sigma^{\prime})}\bigl{)}\Bigr{)}
c0h(x)d(σ,σ)+d(x,x)+|h(x)h(x)|cosd(τ,σ).\displaystyle\leq c_{0}h(x)d(\sigma,\sigma^{\prime})+d(x,x^{\prime})+\frac{|h(x)-h(x^{\prime})|}{-\cos d(\tau,\sigma^{\prime})}.

Since d(x,x)h(x)d(x,x^{\prime})\lesssim h(x^{\prime}), the lemma follows. ∎

Refer to captionAAu×A\langle u\rangle\times Auuvvwweeu,v×e\langle u,v\rangle\times eyyΔ=Δ×y\Delta^{\prime}=\Delta\times yE(Δ)E(\Delta)B(Δ)B(\Delta)
Figure 2. Cells of the “exploded simplex” E(Δ)E(\Delta) are naturally products of cells of Δ\Delta and cells of B(Δ)B(\Delta).

We construct Ω\Omega by piecing together maps of the form ixΔ(Ω(Δ))i_{x_{\Delta}}(\Omega_{\infty}(\Delta)), where Δ\Delta ranges over the simplices of ΔZ\Delta_{Z}. The main problem is that if Δ\Delta^{\prime} is a face of Δ\Delta, the maps ixΔ(Ω(Δ))i_{x_{\Delta}}(\Omega_{\infty}(\Delta)) and ixΔ(Ω(Δ))i_{x_{\Delta^{\prime}}}(\Omega_{\infty}(\Delta^{\prime})) need not agree, since xΔxΔx_{\Delta}\neq x_{\Delta^{\prime}}, so we need to add some “padding” to make these maps agree.

Part of the construction is illustrated in Figure 2: for each simplex Δ\Delta of ΔZ\Delta_{Z}, we “explode” the barycentric subdivision B(Δ)B(\Delta) to get a complex E(Δ)E(\Delta) by inserting a copy Δ\Delta^{\prime} of Δ\Delta in the middle. Each cell in this subdivision is of the form Δ1×Δ2\Delta_{1}\times\Delta_{2}, where Δ1\Delta_{1} is a face of Δ\Delta and Δ2\Delta_{2} is a face of B(Δ)B(\Delta). To be more specific, note that we can label each vertex of B(Δ)B(\Delta) by a face δ\delta of Δ\Delta, and the vertex labels of a simplex δ0,,δk\langle\delta_{0},\dots,\delta_{k}\rangle form a flag δ0δk\delta_{0}\subset\dots\subset\delta_{k}. Then each cell of E(Δ)E(\Delta) is of the form

δ×δ0,,δk,\delta\times\langle\delta_{0},\dots,\delta_{k}\rangle,

for some flag δ0δk\delta_{0}\subset\dots\subset\delta_{k} in Δ\Delta and some face δ\delta of δ0\delta_{0}. The map ρ1:E(Δ)Δ\rho_{1}:E(\Delta)\to\Delta which projects each simplex to its first factor is a continuous map which sends Δ\Delta^{\prime} homeomorphically to Δ\Delta. Likewise, the map ρ2:E(Δ)B(Δ)\rho_{2}:E(\Delta)\to B(\Delta) which projects each cell to its second factor is a continuous map that collapses Δ\Delta^{\prime} to the barycenter of Δ\Delta.

We define a map x:B(Δ)Xx:B(\Delta)\to X on the vertices of B(Δ)B(\Delta) by sending the point δ\langle\delta\rangle to the point xδx_{\delta} for every face δΔ\delta\subset\Delta. We define xx on the rest of B(Δ)B(\Delta) by linear interpolation. That is, if δ0δk\delta_{0}\subset\dots\subset\delta_{k} is a flag of faces of Δ\Delta, then we have xδixδ0+ax_{\delta_{i}}\in x_{\delta_{0}}+a for all ii. Therefore, all the xδix_{\delta_{i}} lie in a common apartment, and we can define xx on δ0,,δk\langle\delta_{0},\dots,\delta_{k}\rangle by linearly interpolating between the xδix_{\delta_{i}}’s. This map has Lipschitz constant diam𝒱(Δ)\lesssim\operatorname{diam}\mathcal{V}(\Delta) on Δ\Delta.

For any cell

σ=δ×δ0,,δk\sigma=\delta\times\langle\delta_{0},\dots,\delta_{k}\rangle

of E(Δ)E(\Delta) and any sσs\in\sigma, let xs=x(ρ2(s))x_{s}=x(\rho_{2}(s)). We have xsxδ+ax_{s}\in x_{\delta}+a and therefore

Ω(δ)Lk(xδ)Lk(xs)\Omega_{\infty}(\delta)\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x_{\delta})\subset\operatorname{\text{Lk}^{\downarrow}_{\infty}}(x_{s})

This means that

ixs(Ω(ρ1(s)))i_{x_{s}}(\Omega_{\infty}(\rho_{1}(s)))

is defined for every sσs\in\sigma, so we define

Ω(s)=ixs(Ω(ρ1(s))).\Omega(s)=i_{x_{s}}(\Omega_{\infty}(\rho_{1}(s))).

Finally, we check that this definition satisfies the conditions of Lemma 3.2. Since xz=zx_{\langle z\rangle}=z for any zZz\in Z, we have Ω(z)=z\Omega(\langle z\rangle)=z, so the first condition is satisfied. Let σ\sigma be a cell of E(Δ)E(\Delta) as above and let s,tσs,t\in\sigma. Let xs=x(ρ2(s))x_{s}=x(\rho_{2}(s)), xt=x(ρ2(t))x_{t}=x(\rho_{2}(t)). By Lemma 4.19, we have

d(Ω(s),Ω(t))cd(xs,xt)+ch(xs)d(Ω(ρ1(s)),Ω(ρ1(t))).d(\Omega(s),\Omega(t))\leq cd(x_{s},x_{t})+ch(x_{s})d(\Omega_{\infty}(\rho_{1}(s)),\Omega_{\infty}(\rho_{1}(t))).

Since xΔxδi+𝔞x_{\Delta}\in x_{\delta_{i}}+\mathfrak{a} for each i=1,ki=1,\dots k, we have xΔxs+𝔞x_{\Delta}\in x_{s}+\mathfrak{a} and thus h(xs)diam𝒱(Δ)h(x_{s})\lesssim\operatorname{diam}\mathcal{V}(\Delta). Since ρ1\rho_{1}, ρ2\rho_{2}, and Ω\Omega_{\infty} are Lipschitz with constants depending only on XX and Lip(x|Δ)diam𝒱(Δ)\operatorname{Lip}(x|_{\Delta})\lesssim\operatorname{diam}\mathcal{V}(\Delta), each term in the inequality above is diam𝒱(Δ)d(s,t)\lesssim\operatorname{diam}\mathcal{V}(\Delta)d(s,t). Therefore,

Lip(Ω|Δ)diam𝒱(Δ)\operatorname{Lip}(\Omega_{\infty}|_{\Delta})\lesssim\operatorname{diam}\mathcal{V}(\Delta)

for every simplex ΔΔZ(n1)\Delta\subset\Delta_{Z}^{(n-1)}, as desired.

This proves Lemma 4.2.

References

  • [AB08] P. Abramenko and K. S. Brown. Buildings, volume 248 of Graduate Texts in Mathematics. Springer, New York, 2008. Theory and applications.
  • [ABD+12] A. Abrams, N. Brady, P. Dani, M. Duchin, and R. Young. Pushing fillings in right-angled artin groups. Journal of the London Mathematical Society, 2012.
  • [Abr96] P. Abramenko. Twin buildings and applications to S-arithmetic groups, volume 1641 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1996.
  • [AK00] L. Ambrosio and B. Kirchheim. Currents in metric spaces. Acta Math., 185(1):1–80, 2000.
  • [BEW] M. Bestvina, A. Eskin, and K. Wortman. Filling boundaries of coarse manifolds in semisimple and solvable arithmetic groups. arXiv:1106.0162.
  • [BW07] K.-U. Bux and K. Wortman. Finiteness properties of arithmetic groups over function fields. Invent. Math., 167(2):355–378, 2007.
  • [BW11] K.-U. Bux and K. Wortman. Connectivity properties of horospheres in Euclidean buildings and applications to finiteness properties of discrete groups. Invent. Math., 185(2):395–419, 2011.
  • [dCT10] Y. de Cornulier and R. Tessera. Metabelian groups with quadratic Dehn function and Baumslag-Solitar groups. Confluentes Math., 2(4):431–443, 2010.
  • [Dru04] C. Druţu. Filling in solvable groups and in lattices in semisimple groups. Topology, 43(5):983–1033, 2004.
  • [ECH+92] D. B. A. Epstein, J. W. Cannon, D. F. Holt, S. V. F. Levy, M. S. Paterson, and W. P. Thurston. Word processing in groups. Jones and Bartlett Publishers, Boston, MA, 1992.
  • [Ger93] S. M. Gersten. Isoperimetric and isodiametric functions of finite presentations. In Geometric group theory, Vol. 1 (Sussex, 1991), volume 181 of London Math. Soc. Lecture Note Ser., pages 79–96. Cambridge Univ. Press, Cambridge, 1993.
  • [Gro] C. Groft. Generalized Dehn functions II. arXiv:0901.2317.
  • [Gro83] M. Gromov. Filling Riemannian manifolds. J. Differential Geom., 18(1):1–147, 1983.
  • [Gro93] M. Gromov. Asymptotic invariants of infinite groups. In Geometric group theory, Vol. 2 (Sussex, 1991), volume 182 of London Math. Soc. Lecture Note Ser., pages 1–295. Cambridge Univ. Press, Cambridge, 1993.
  • [Gro96] M. Gromov. Carnot-Carathéodory spaces seen from within. In Sub-Riemannian geometry, volume 144 of Progr. Math., pages 79–323. Birkhäuser, Basel, 1996.
  • [KL97] B. Kleiner and B. Leeb. Rigidity of quasi-isometries for symmetric spaces and Euclidean buildings. Inst. Hautes Études Sci. Publ. Math., (86):115–197 (1998), 1997.
  • [LMR00] A. Lubotzky, S. Mozes, and M. S. Raghunathan. The word and Riemannian metrics on lattices of semisimple groups. Inst. Hautes Études Sci. Publ. Math., (91):5–53 (2001), 2000.
  • [LP96] E. Leuzinger and C. Pittet. Isoperimetric inequalities for lattices in semisimple Lie groups of rank 22. Geom. Funct. Anal., 6(3):489–511, 1996.
  • [LS05] U. Lang and T. Schlichenmaier. Nagata dimension, quasisymmetric embeddings, and Lipschitz extensions. Int. Math. Res. Not., (58):3625–3655, 2005.
  • [Pit95] C. Pittet. Hilbert modular groups and isoperimetric inequalities. In Combinatorial and geometric group theory (Edinburgh, 1993), volume 204 of London Math. Soc. Lecture Note Ser., pages 259–268. Cambridge Univ. Press, Cambridge, 1995.
  • [Sch] B. Schulz. Spherical subcomplexes of spherical buildings. arXiv:1007.2407.
  • [Wen08] S. Wenger. A short proof of Gromov’s filling inequality. Proc. Amer. Math. Soc., 136(8):2937–2941, 2008.
  • [Whi84] B. White. Mappings that minimize area in their homotopy classes. J. Differential Geom., 20(2):433–446, 1984.
  • [You] R. Young. The Dehn function of SL(n;){\rm SL}(n;\mathbb{Z}). arXiv:0912.2697.