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

Sub-Finsler horofunction boundaries of the Heisenberg group

Nate Fisher Department of Mathematics, 503 Boston Ave, Medford, MA [email protected]  and  Sebastiano Nicolussi Golo Department of Mathematics and Statistics, 40014 University of Jyväskylä, Finland [email protected]
Abstract.

We give a complete analytic and geometric description of the horofunction boundary for polygonal sub-Finsler metrics—that is, those that arise as asymptotic cones of word metrics—on the Heisenberg group. We develop theory for the more general case of horofunction boundaries in homogeneous groups by connecting horofunctions to Pansu derivatives of the distance function.

Key words and phrases:
Horoboundary, sub-Finsler distance, homogeneous group, Heisenberg group
2010 Mathematics Subject Classification:
20F69,53C23,53C17
S.N.G has been supported by the University of Padova STARS Project “Sub-Riemannian Geometry and Geometric Measure Theory Issues: Old and New”; by the INdAM – GNAMPA Project 2019 “Rectifiability in Carnot groups”; and by the Marie Curie Actions-Initial Training Network “Metric Analysis For Emergent Technologies (MAnET)” (n. 607643).

1. Introduction

1.1. Describing the horofunction boundary

The study of boundaries of metric spaces has a rich history and has been fundamental in building bridges between the fields of algebra, topology, geometry, and dynamical systems. Understanding the boundary was essential in the proof of Mostow’s rigidity theorem for closed hyperbolic manifolds, and boundaries have also been used to classify isometries of metric spaces, to understand algebraic splittings of groups, and to study the asymptotic behavior of random walks.

The simplest and most classical setting for horofunctions is in the study of isometries of the hyperbolic plane. There, the isometry group splits and induces a geodesic flow and a horocycle flow on the tangent bundle; horocycles, or orbits of the horocycle flow, are level sets of horofunctions. The notion has since been abstracted by Busemann, generalized by Gromov, and used by Rieffel, Karlsson–Ledrappier, and many others to derive results in various fields. The horofunction boundary is obtained by embedding a metric space XX into the space of continuous real-valued functions on XX via the metric, as we will define below.

In this paper, we develop tools to study the horofunction boundary of homogeneous groups, in particular the real Heisenberg group \mathbb{H}. The horofunction boundary of the Heisenberg group has been the subject of study in several publications. Klein and Nicas described the boundary of \mathbb{H} for the Korányi and sub-Riemannian metrics [12, 13], while several others have studied the boundaries of discrete word metrics in the integer Heisenberg group [24, 1]. In this paper, we aim to understand the horofunction boundary of the real Heisenberg group \mathbb{H} for a family of polygonal sub-Finsler metrics which arise as the asymptotic cones of the integer Heisenberg group for different word metrics [21].

While horofunction boundaries are not (yet) used as widely as visual boundaries or Poisson boundaries, they admit a theory which is useful across several fields including geometry, analysis, and dynamical systems. Whether it is classifying Busemann function, giving explicit formulas for the horofunctions, describing the topology of the boundary, or studying the action of isometries on the boundary, what it means to understand or to describe a horofunction boundary varies significantly between works.

In this paper, as is done in for the \ell^{\infty} metric on n\mathbb{R}^{n} in [5], we hope to combine these analytic, topological, and dynamical descriptions while also introducing a more geometric approach. In particular, we want to associate a “direction” to every horofunction as well as a geometric condition for a sequence of points to induce a horofunction. In some settings, the horofunction boundary is made up entirely of limit points induced by geodesic rays—or in other words, every horofunction is a Busemann function. It is known that in CAT(0) spaces [2] as well as in polyhedral normed vector spaces [11], the horofunction boundary is composed only of Busemann functions. This connection between horofunctions and geodesic rays provides a natural notion of directionality to the horofunction boundary, which is not present in settings of mixed curvature, as described in [16]. For the model we develop in homogeneous metrics, sequences converging to a horofunction can often be dilated back to a well-defined point on the unit sphere, which we can then regard as a direction. In these sub-Finsler metrics, there are many directions with no infinite geodesics at all, so this provides one of the motivating senses in which the horofunction boundary is a better choice to capture the geometry and dynamics in nilpotent groups.

1.2. Outline of paper

For any homogeneous group, we convert the problem of describing the horofunction boundary to a study of directional derivatives, i.e., Pansu derivatives, of the distance function. It suffices to understand Pansu derivatives on the unit sphere. Therefore, in any homogeneous group where the unit sphere is understood, our method allows a description of the horofunction boundary.

Pansu-differentiable points on the sphere (i.e., points pp at which distance to the origin has a well defined Pansu derivative) can be thought of as directions of horofunctions. Not all horofunctions are directional; the rest are blow-ups of non-differentiable points. Background on homogeneous groups, Pansu derivatives, and horofunctions is provided in §2. We use Kuratowski limits—a notion of set convergence in a metric space—to define the blow-up of a function in §3.

In the remainder of the paper, we focus on the Heisenberg group \mathbb{H}. For sub-Riemannian metrics on \mathbb{H}, Klein–Nicas showed that the horofunction boundary is a topological disk [13]. In Theorem 4.1 of §4 we show that an analogous disk belongs to the boundary for the larger class of sub-Finsler metrics, but is a proper subset in many cases.

Our main theorem (Theorem 5.4 in §5) describes the horoboundary of polygonal sub-Finsler metrics on \mathbb{H} in terms of blow-ups. From this, we are able to give explicit expressions for the horofunctions, to describe the topology of the boundary, and to identify Busemann points.

This description is extremely explicit and allows us to visualize the horofunction boundary and to understand it geometrically. We get a correspondence between “directions” on the sphere and functions in the boundary, as indicated in Figure 1. This description allows us to realize the horofunction boundary as a kind of dual to the unit sphere, generalizing previous observations for normed vector spaces and for the sub-Riemannian metric on \mathbb{H} [9, 5, 11, 23, 13].

Refer to captionRefer to captionRefer to captionRefer to caption(2,d𝖤𝗎𝖼𝗅)(\mathbb{R}^{2},d_{\sf Eucl})(2,d𝗁𝖾𝗑)(\mathbb{R}^{2},d_{\sf hex})(,d𝗌𝗎𝖻𝖱𝗂𝖾𝗆)(\mathbb{H},d_{\sf subRiem})(,d𝗌𝗎𝖻𝖥𝗂𝗇𝗌)(\mathbb{H},d_{\sf subFins})SphereSphereSphereSphereBoundaryBoundaryBoundaryBoundary
Figure 1. The duality between unit spheres and horofunction boundaries for various metric spaces, where colors indicate a correspondence between directions on the spheres and points in the boundary. Note that in both the round and hexagonal cases, the 2D spheres and boundaries embed in the Heisenberg spheres (along the equators) and boundaries.

Finally, using our description of the boundary, we also study the group action on the boundary in §6, generalizing results of Walsh and Bader–Finkelshtein [24, 1].

Acknowledgements

The authors would like to thank Moon Duchin for suggesting the problem and bringing us together to work on this project. We also appreciate the fruitful discussions we have had with Enrico Le Donne, Sunrose Shrestha, and Anders Karlsson. Finally, we thank Linus Kramer for pointing out to us a common mistake in the definition of horoboundary that we had repeated, see Section 2.5.

2. Preliminaries on homogeneous groups and horofunctions

We begin with a brief introduction to graded Lie groups, homogeneous metrics, Pansu derivatives, and horofunctions. For a survey on graded Lie groups and homogeneous metrics, we refer the interested reader to [14].

2.1. Graded Lie groups

Let VV be a real vector space with finite dimension and [,]:V×VV[\cdot,\cdot]:V\times V\to V be the Lie bracket of a Lie algebra 𝔤=(V,[,])\mathfrak{g}=(V,[\cdot,\cdot]). We say that 𝔤\mathfrak{g} is graded if subspaces V1,,VsV_{1},\dots,V_{s} are fixed so that V=V1VsV=V_{1}\oplus\dots\oplus V_{s} and [Vi,Vj]:=span{[v,w]:vVi,wVj}Vi+j[V_{i},V_{j}]:=\mathrm{span}\{[v,w]:v\in V_{i},\ w\in V_{j}\}\subset V_{i+j} for all i,j{1,,s}i,j\in\{1,\dots,s\}, where Vk={0}V_{k}=\{0\} if k>sk>s. Graded Lie algebras are nilpotent. A graded Lie algebra is stratified of step ss if equality [V1,Vj]=Vj+1[V_{1},V_{j}]=V_{j+1} holds and Vs{0}V_{s}\neq\{0\}. Our main object of study are stratified Lie algebras, but we will often work with subspaces that are only graded Lie algebras.

On the vector space VV we define a group operation via the Baker–Campbell–Hausdorff formula

pq\displaystyle pq :=n=1(1)n1n{sj+rj>0:j=1n}[pr1qs1pr2qs2prnqsn]j=1n(rj+sj)i=1nri!si!\displaystyle:=\sum_{n=1}^{\infty}\frac{(-1)^{n-1}}{n}\sum_{\{s_{j}+r_{j}>0:j=1\dots n\}}\frac{[p^{r_{1}}q^{s_{1}}p^{r_{2}}q^{s_{2}}\cdots p^{r_{n}}q^{s_{n}}]}{\sum_{j=1}^{n}(r_{j}+s_{j})\prod_{i=1}^{n}r_{i}!s_{i}!}
=p+q+12[p,q]+,\displaystyle=p+q+\frac{1}{2}[p,q]+\dots,

where

[pr1qs1pr2qs2prnqsn]=[p,[p,,r1 times[q,[q,,s1 times[p,]]]]].[p^{r_{1}}q^{s_{1}}p^{r_{2}}q^{s_{2}}\cdots p^{r_{n}}q^{s_{n}}]=\underbrace{[p,[p,\dots,}_{r_{1}\text{ times}}\underbrace{[q,[q,\dots,}_{s_{1}\text{ times}}\underbrace{[p,\dots}_{\dots}]\dots]]\dots]].

The sum in the formula above is finite because 𝔤\mathfrak{g} is nilpotent. The resulting Lie group, which we denote by 𝔾\mathbb{G}, is nilpotent and simply connected; we will call it graded group or stratified group, depending on the type of grading of the Lie algebra. The identification 𝔾=V=𝔤\mathbb{G}=V=\mathfrak{g} corresponds to the identification between Lie algebra and Lie group via the exponential map exp:𝔤𝔾\exp:\mathfrak{g}\to\mathbb{G}. Notice that p1=pp^{-1}=-p for every p𝔾p\in\mathbb{G} and that 0 is the neutral element of 𝔾\mathbb{G}.

If 𝔤\mathfrak{g}^{\prime} is another graded Lie algebra with underlying vector space VV^{\prime} and Lie group 𝔾\mathbb{G}^{\prime}, then, with the same identifications as above, a map VVV\to V^{\prime} is a Lie algebra morphism if and only if it is a Lie group morphism, and all such maps are linear. In particular, we denote by Homh(𝔾;𝔾)\operatorname{Hom}_{h}(\mathbb{G};\mathbb{G}^{\prime}) the space of all homogeneous morphisms from 𝔾\mathbb{G} to 𝔾\mathbb{G}^{\prime}, that is, all linear maps VVV\to V^{\prime} that are Lie algebra morphisms (equivalently, Lie group morphisms) and that map VjV_{j} to VjV_{j}^{\prime}. If 𝔤\mathfrak{g} is stratified, then homogeneous morphisms are uniquely determined by their restriction to V1V_{1}.

For λ>0\lambda>0, define the dilations as the maps δλ:VV\delta_{\lambda}:V\to V such that δλv=λjv\delta_{\lambda}v=\lambda^{j}v for vVjv\in V_{j}. Notice that δλδμ=δλμ\delta_{\lambda}\delta_{\mu}=\delta_{\lambda\mu} and that δλHomh(𝔾;𝔾)\delta_{\lambda}\in\operatorname{Hom}_{h}(\mathbb{G};\mathbb{G}), for all λ,μ>0\lambda,\mu>0. Notice also that a Lie group morphism F:𝔾𝔾F:\mathbb{G}\to\mathbb{G}^{\prime} is homogeneous if and only if Fδλ=δλFF\circ\delta_{\lambda}=\delta_{\lambda}^{\prime}\circ F for all λ>0\lambda>0, where δλ\delta^{\prime}_{\lambda} denotes the dilations in 𝔾\mathbb{G}^{\prime}. We say that a subset MM of VV is homogeneous if δλ(M)=M\delta_{\lambda}(M)=M for all λ>0\lambda>0.

A homogeneous distance on 𝔾\mathbb{G} is a distance function dd that is left-invariant and 1-homogeneous with respect to dilations, i.e.,

  1. (i)

    d(gx,gy)=d(x,y)d(gx,gy)=d(x,y) for all g,x,y𝔾g,x,y\in\mathbb{G};

  2. (ii)

    d(δλx,δλy)=λd(x,y)d(\delta_{\lambda}x,\delta_{\lambda}y)=\lambda d(x,y) for all x,y𝔾x,y\in\mathbb{G} and all λ>0\lambda>0.

When a stratified group 𝔾\mathbb{G} is endowed with a homogeneous distance dd, we call the metric Lie group (𝔾,d)(\mathbb{G},d) a Carnot group. Homogeneous distances induce the topology of 𝔾\mathbb{G}, see [17, Proposition 2.26], and are biLipschitz equivalent to each other. Every homogeneous distance defines a homogeneous norm de():𝔾[0,),de(p)=d(e,p)d_{e}(\cdot):\mathbb{G}\to[0,\infty),d_{e}(p)=d(e,p), where ee is the neutral element of 𝔾\mathbb{G}. We denote by |||\cdot| the Euclidean norm in \mathbb{R}^{\ell}.

2.2. Pansu derivatives

Let 𝔾\mathbb{G} and 𝔾\mathbb{G}^{\prime} be two Carnot groups with homogeneous metrics dd and dd^{\prime}, respectively, and let Ω𝔾\Omega\subset\mathbb{G} open. A function f:Ω𝔾f:\Omega\to\mathbb{G}^{\prime} is Pansu differentiable at pΩp\in\Omega if there is LHomh(𝔾;𝔾)L\in\operatorname{Hom}_{h}(\mathbb{G};\mathbb{G}^{\prime}) such that

limxpd(f(p)1f(x),L(p1x))d(p,x)=0.\lim_{x\to p}\frac{d^{\prime}(f(p)^{-1}f(x),L(p^{-1}x))}{d(p,x)}=0.

The map LL is called Pansu derivative of ff at pp and it is denoted by

P

​​D
f(p)
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}f(p)
or

P

​​D
f
|
p
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}f|_{p}
. A map f:Ω𝔾f:\Omega\to\mathbb{G}^{\prime} is of class CH1C^{1}_{H} if ff is Pansu differentiable at all points of Ω\Omega and the Pansu derivative p

P

​​D
F
|
p
p\mapsto\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F|_{p}
is continuous. We denote by CH1(Ω;𝔾)C^{1}_{H}(\Omega;\mathbb{G}^{\prime}) the space of all maps from Ω\Omega to 𝔾\mathbb{G}^{\prime} of class CH1C^{1}_{H}.

A function f:Ω𝔾f:\Omega\to\mathbb{G}^{\prime} is strictly Pansu differentiable at pΩp\in\Omega if there is LHomh(𝔾;𝔾)L\in\operatorname{Hom}_{h}(\mathbb{G};\mathbb{G}^{\prime}) such that

limϵ0sup{d(f(y)1f(x),L(y1x))d(x,y):x,yBd(p,ϵ),xy}=0,\lim_{\epsilon\to 0}\ \sup\left\{\frac{d^{\prime}(f(y)^{-1}f(x),L(y^{-1}x))}{d(x,y)}:x,y\in B_{d}(p,\epsilon),\ x\neq y\right\}=0,

where Bd(p,ϵ)B_{d}(p,\epsilon) is the open ϵ\epsilon-ball centered at pp. Clearly, in this case ff is Pansu differentiable at pp and L=

P

​​D
F
|
p
L=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F|_{p}
. Moreover, as shown in [10, Proposition 2.4 and Lemma 2.5], a function f:Ω𝔾f:\Omega\to\mathbb{G}^{\prime} is of class CH1C^{1}_{H} on Ω\Omega if and only if ff is strictly Pansu differentiable at all points in Ω\Omega. If fCH1(Ω;𝔾)f\in C^{1}_{H}(\Omega;\mathbb{G}^{\prime}), then f:(Ω,d)(𝔾,d)f:(\Omega,d)\to(\mathbb{G}^{\prime},d^{\prime}) is locally Lipschitz.

2.3. Sub-Finsler metrics

Let 𝔾\mathbb{G} be a stratified group and \|\cdot\| a norm on the first layer V1Te𝔾V_{1}\subset T_{e}\mathbb{G} of the stratification. Using left-translations, we extend the norm \|\cdot\| to the sub-bundle ΔT𝔾\Delta\subset T\mathbb{G} of left-translates of V1V_{1}. We call a curve in 𝔾\mathbb{G} admissible if it is tangent to Δ\Delta almost everywhere, and using the norm \|\cdot\| we can measure the length of any admissible curve. A classical result tells us that in a stratified group, where V1V_{1} bracket-generates the whole Lie algebra, any two points in 𝔾\mathbb{G} are connected by an admissible curve. We then define a Carnot-Carathéodory length metric by

d(p,q)=infγ{abγ(t)𝑑t},d(p,q)=\inf_{\gamma}\left\{\int_{a}^{b}\lVert\gamma^{\prime}(t)\rVert dt\right\},

where the infimum is taken over all admissible γ\gamma connecting pp to qq.

Proposition 2.1 (Eikonal equation).

If dd is a homogeneous distance on 𝔾\mathbb{G}, then de:xd(e,x)d_{e}:x\mapsto d(e,x) is Pansu differentiable almost everywhere. Moreover, if dd is sub-Finsler with norm \|\cdot\|, then

(1)

P

​​D
de|p:=sup{|

P

​​D
de|pv|:d(e,v)1}
=1for a.e. p𝔾.
\|\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}\|:=\sup\{|\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}v|:d(e,v)\leq 1\}=1\qquad\text{for a.e.\leavevmode\nobreak\ }p\in\mathbb{G}.
Proof.

Since ded_{e} is 1-Lipschitz, then it is Pansu differentiable almost everywhere by the Pansu–Rademacher Theorem [22, Theorem 2] and

P

​​D
de
1
\|\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}\|\leq 1
. To prove (1), let p𝔾p\in\mathbb{G} be a point at which ded_{e} is Pansu differentiable, and let γ:[0,T]𝔾\gamma:[0,T]\to\mathbb{G} be a length minimizing curve parametrized by arc-length from ee to pp. Since, for tTt\neq T, we have

d(e,δ1/|tT|(γ(T)1γ(t)))=1,d\left(e,\delta_{1/|t-T|}\left(\gamma(T)^{-1}\gamma(t)\right)\right)=1,

then there is a sequence tnTt_{n}\to T so that limnδ1/|tnT|(γ(T)1γ(tn))=v\lim_{n\to\infty}\delta_{1/|t_{n}-T|}\left(\gamma(T)^{-1}\gamma(t_{n})\right)=v exists. It follows that

1=limn|de(γ(tn))de(γ(T))||tnT|=|

P

​​D
de|p[v]|,
1=\lim_{n\to\infty}\frac{|d_{e}(\gamma(t_{n}))-d_{e}(\gamma(T))|}{|t_{n}-T|}=|\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}[v]|,

and we conclude that

P

​​D
de|p=1
\|\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}\|=1
. ∎

2.4. The Heisenberg group

The Heisenberg group \mathbb{H} is the simply connected Lie group whose Lie algebra 𝔥\mathfrak{h} is generated by three vectors XX, YY, and ZZ, with the only nontrivial Lie bracket [X,Y]=Z[X,Y]=Z. The stratification is given by V1=span{X,Y}V_{1}=\mathrm{span}\{X,Y\} and V2=span{Z}V_{2}=\mathrm{span}\{Z\}. Via the exponential map and the above basis for 𝔥\mathfrak{h}, the Heisenberg group can be coordinatized as 3\mathbb{R}^{3} with the following group multiplication:

(x,y,z)(x,y,z)=(x+x,y+y,z+z+12(xyxy)).(x,y,z)(x^{\prime},y^{\prime},z^{\prime})=(x+x^{\prime},y+y^{\prime},z+z^{\prime}+\frac{1}{2}(xy^{\prime}-x^{\prime}y)).

Under this group operation, the generating vectors in the Lie algebra correspond to the left-invariant vector fields

X=x12yz,Y=y+12xz,Z=z.X=\partial_{x}-\frac{1}{2}y\partial_{z},\quad Y=\partial_{y}+\frac{1}{2}x\partial z,\quad Z=\partial_{z}.

It will sometimes be convenient to coordinatize \mathbb{H} as 2×\mathbb{R}^{2}\times\mathbb{R}, in which case the group operation can be written

(v,t)(w,s)=(v+w,t+s+12ω(v,w)),(v,t)(w,s)=(v+w,t+s+\frac{1}{2}\omega(v,w)),

where ω\omega is the standard symplectic form on the plane, ω((x,y),(x,y))=xyxy\omega((x,y),(x^{\prime},y^{\prime}))=xy^{\prime}-x^{\prime}y.

Denote by Δ\Delta the horizontal distribution, the sub-bundle (or plane field) generated by the vector fields XX and YY. A curve is admissible if its derivative belongs to Δ\Delta.

Let π:2\pi:\mathbb{H}\to\mathbb{R}^{2} be the projection of a point to its horizontal components, π(x,y,z)=(x,y)\pi(x,y,z)=(x,y), which is a group morphism.

Given a path γ:[0,T]2\gamma:[0,T]\to\mathbb{R}^{2} and an initial height z0z_{0}, there exists a unique lift to an admissible path γ^\hat{\gamma} in \mathbb{H} such that γ^\hat{\gamma} has height z0z_{0} at time zero and π(γ^)=γ\pi(\hat{\gamma})=\gamma. Using Green’s theorem and applying an elementary observation, we have that the third component of γ^\hat{\gamma},

γ3(t)=z0+120t(γ1γ2γ2γ1)(s)ds,\gamma_{3}(t)=z_{0}+\frac{1}{2}\int_{0}^{t}(\gamma_{1}\gamma_{2}^{\prime}-\gamma_{2}\gamma_{1}^{\prime})(s)\,\mathrm{d}s,

is given by the sum of z0z_{0} and the balayage area of γ\gamma, i.e., the signed area enclosed by γ\gamma.

Let dd be the sub-Finsler metric on \mathbb{H} induced by a norm \|\cdot\| on 2\mathbb{R}^{2} with unit disk QQ. The length in (,d)(\mathbb{H},d) of an admissible curve γ^\hat{\gamma} is equal to the length in (2,)(\mathbb{R}^{2},\|\cdot\|) of the projected curve π(γ^)\pi(\hat{\gamma}). A well-known result is that geodesics in sub-Finsler metrics are lifts of solutions to the Dido problem with respect to \|\cdot\|; that is, geodesics are lifts of arcs which trace the perimeter of the isoperimetrix II for the given norm.

2.5. Horoboundary of a metric space

Let (X,d)(X,d) be a metric space and 𝒞(X)\mathscr{C}(X) the space of continuous functions XX\to\mathbb{R} endowed with the topology of the uniform convergence on compact sets. The map ι:X𝒞(X)\iota:X\hookrightarrow\mathscr{C}(X), (ι(x))(y):=d(x,y)(\iota(x))(y):=d(x,y), is an embedding, i.e., a homeomorphism onto its image.

Let 𝒞(X)/\mathscr{C}(X)/\mathbb{R} be the topological quotient of 𝒞(X)\mathscr{C}(X) with kernel the constant functions, i.e., for every f,g𝒞(X)f,g\in\mathscr{C}(X) we set the equivalence relation fgfgf\sim g\Leftrightarrow f-g is constant. For oXo\in X, we set

𝒞(X)o:={f𝒞(X):f(o)=0}.\mathscr{C}(X)_{o}:=\{f\in\mathscr{C}(X):f(o)=0\}.

Then the restriction 𝒞(X)o𝒞(X)/\mathscr{C}(X)_{o}\to\mathscr{C}(X)/\mathbb{R} of the quotient map is an isomorphism of topological vector spaces, for each oXo\in X. Indeed, one easily checks that it is both injective and surjective, and that its inverse map is [f]ff(o)[f]\mapsto f-f(o), where [f]𝒞(X)/[f]\in\mathscr{C}(X)/\mathbb{R} is the class of equivalence of f𝒞(X)f\in\mathscr{C}(X), is continuous.

The map ι^:XC(X)𝒞(X)/\hat{\iota}:X\hookrightarrow C(X)\to\mathscr{C}(X)/\mathbb{R} is injective. Indeed, if x,xXx,x^{\prime}\in X are such that ι(x)(z)ι(x)(z)\iota(x)(z)-\iota(x^{\prime})(z) is constant for all zXz\in X, then taking z=xz=x and then z=xz=x^{\prime} in turn tells us that c=d(x,x)=d(x,x)c=d(x,x^{\prime})=-d(x^{\prime},x). Hence c=0c=0 and x=xx=x^{\prime}.

Moreover, ι^\hat{\iota} is continuous, but it does not need to be an embedding, as we learned from [4, Proposition 4.5]. In the following lemma, which is a generalization of  [4, Remark 4.3], we show that ι^\hat{\iota} is an embedding under mild conditions on the distance function.

Lemma 2.2.

Let (X,d)(X,d) be a proper metric space with the following property:

(2) pX0<r<sxXB(p,s)zB(p,s)B(p,r) s.t. d(x,z)d(x,p).\forall p\in X\,\exists 0<r<s\,\forall x\in X\setminus B(p,s)\,\exists z\in B(p,s)\setminus B(p,r)\text{ s.t. }d(x,z)\leq d(x,p).

Then the map ι^:X𝒞(X)/\hat{\iota}:X\hookrightarrow\mathscr{C}(X)/\mathbb{R} is an embedding.

In particular, any proper metric space with path connected balls satisfy (2) with r=1r=1 and s=2s=2. And so do homogeneous distances on graded groups.

Proof.

We need to show that ι^\hat{\iota} maps closed sets to closed subsets of ι^(X)\hat{\iota}(X). Let AXA\subset X closed and pXAp\in X\setminus A: we claim that ι^(p)cl(ι^(A))\hat{\iota}(p)\notin\text{cl}(\hat{\iota}(A)).

Using the isomorphism 𝒞(X)p𝒞(X)/\mathscr{C}(X)_{p}\to\mathscr{C}(X)/\mathbb{R}, we can prove the claim for the map ι^p:X𝒞(X)p\hat{\iota}_{p}:X\to\mathscr{C}(X)_{p}, ι^p(x)=d(x,)d(x,p)\hat{\iota}_{p}(x)=d(x,\cdot)-d(x,p). Let 0<r<s0<r<s as in (2) for this pp, and let ϵ>0\epsilon>0 be such that B(p,2ϵ)A=B(p,2\epsilon)\cap A=\emptyset. We show that, for every xAx\in A,

(3) supzB¯(p,s)|ι^p(p)(z)ι^p(x)(z)|min{4ϵ,r}.\sup_{z\in\bar{B}(p,s)}|\hat{\iota}_{p}(p)(z)-\hat{\iota}_{p}(x)(z)|\geq\min\{4\epsilon,r\}.

Fix xAx\in A. First, if d(p,x)sd(p,x)\leq s, then

ι^p(p)(x)ι^p(x)(x)=2d(p,x)4ϵ.\hat{\iota}_{p}(p)(x)-\hat{\iota}_{p}(x)(x)=2d(p,x)\geq 4\epsilon.

Second, if d(p,x)>sd(p,x)>s, then let zz as in (2), so that

ι^p(p)(z)ι^p(x)(z)=d(p,z)(d(x,z)d(x,p))d(p,z)r.\hat{\iota}_{p}(p)(z)-\hat{\iota}_{p}(x)(z)=d(p,z)-(d(x,z)-d(x,p))\geq d(p,z)\geq r.

Thus (3). We conclude from (3) and the compactness of B¯(p,s)\bar{B}(p,s) that ι^p(p)cl(ι^p(A))\hat{\iota}_{p}(p)\notin\text{cl}(\hat{\iota}_{p}(A)) and thus ι^(p)cl(ι^(A))\hat{\iota}(p)\notin\text{cl}(\hat{\iota}(A)), where cl()\text{cl}(\cdot) denotes the topological closure. Hence, ι^(A)=cl(ι^(A)ι^(X))\hat{\iota}(A)=\text{cl}(\hat{\iota}(A)\cap\hat{\iota}(X)), i.e., ι^(A)\hat{\iota}(A) is a closed subset of ι^(X)\hat{\iota}(X). This completes the proof of the first part of the lemma.

For the second part, let (X,d)(X,d) be a proper metric space with path connected balls. Set r=1r=1 and s=2s=2, let pXp\in X and xXx\in X with d(p,x)sd(p,x)\geq s. Since B¯(x,d(p,x))\bar{B}(x,d(p,x)) is path connected, there is a continuous curve γ:[0,1]B¯(x,d(p,x))\gamma:[0,1]\to\bar{B}(x,d(p,x)) with γ(0)=x\gamma(0)=x and γ(1)=p\gamma(1)=p. Since a(t):=td(p,γ(t))a(t):=t\mapsto d(p,\gamma(t)) is continuous, a(0)sa(0)\geq s and a(1)=0a(1)=0, then there is t0t_{0} with d(p,γ(t0))[r,s]d(p,\gamma(t_{0}))\in[r,s]. We conclude that (2) holds with z=γ(t0)z=\gamma(t_{0}).

Notice that homogeneous distances on graded groups satisfy the above connectedness condition. ∎

Define the horoboundary of (X,d)(X,d) as

hX:=cl(ι^(X))ι^(X)𝒞(X)/,\partial_{h}X:=cl(\hat{\iota}(X))\setminus\hat{\iota}(X)\subset\mathscr{C}(X)/\mathbb{R},

where cl(ι^(X))cl(\hat{\iota}(X)) is the topological closure. Another description of the horoboundary is possible,

as we identify hX\partial_{h}X with a subset of 𝒞(X)o\mathscr{C}(X)_{o} for some oXo\in X. More explicitly: f𝒞(X)of\in\mathscr{C}(X)_{o} belongs to hX\partial_{h}X if and only if there is a sequence pnXp_{n}\in X such that pnp_{n}\to\infty (i.e., for every compact KXK\subset X there is NN\in\mathbb{N} such that pnKp_{n}\notin K for all n>Nn>N) and the sequence of functions fn𝒞(X)of_{n}\in\mathscr{C}(X)_{o},

(4) fn(x):=d(pn,x)d(pn,o),f_{n}(x):=d(p_{n},x)-d(p_{n},o),

converge uniformly on compact sets to ff.

If γ:[0,)X\gamma:[0,\infty)\to X is a geodesic ray, one can check that limtι^(γ(t))\lim_{t\to\infty}\hat{\iota}(\gamma(t)) exists, and the geodesic ray converges to a horofunction. Indeed, one can check that for each xx in a compact set KK, {d(γ(t),x)d(γ(t),γ(0))}\{d(\gamma(t),x)-d(\gamma(t),\gamma(0))\} is non-increasing and bounded below. These horofunctions which are the limits of geodesic rays, Busemann functions, have been widely studied and inspired the definition of general horofunctions.

2.6. Horofunctions and the Pansu derivative

On homogeneous groups, we observe a fundamental connection between horofunctions and Pansu derivatives of the function de:xd(e,x)d_{e}:x\mapsto d(e,x).

Let dd be a homogeneous metric on 𝔾\mathbb{G} with unit ball BB and unit sphere B\partial B. Again, we denote by ee the neutral element of 𝔾\mathbb{G} and by ded_{e} the function xd(e,x)x\mapsto d(e,x).

Lemma 2.3.

Let dd be a homogeneous metric on 𝔾\mathbb{G}. If fh(𝔾,d)f\in\partial_{h}(\mathbb{G},d), then there is a sequence (pn,ϵn)B×(0,+)(p_{n},\epsilon_{n})\in\partial B\times(0,+\infty) such that pnpBp_{n}\to p\in\partial B, ϵn0\epsilon_{n}\to 0 and

(5) f(x)=limnde(pnδϵnx)de(pn)ϵn,locally uniformly in x𝔾.f(x)=\lim_{n\to\infty}\frac{d_{e}(p_{n}\delta_{\epsilon_{n}}x)-d_{e}(p_{n})}{\epsilon_{n}},\quad\text{locally uniformly in $x\in\mathbb{G}$.}

On the other hand, if (pn,ϵn)𝔾×(0,+)(p_{n},\epsilon_{n})\in\mathbb{G}\times(0,+\infty) such that pnpBp_{n}\to p\in\partial B, ϵn0\epsilon_{n}\to 0 and f:𝔾f:\mathbb{G}\to\mathbb{R} is the locally uniform limit (5), then fh(𝔾,d)f\in\partial_{h}(\mathbb{G},d).

The horofunction ff is limit of the sequence of points

(6) qn=δ1/ϵnpn1.q_{n}=\delta_{1/\epsilon_{n}}p_{n}^{-1}.

Moreover, if ded_{e} is strictly Pansu differentiable at pp, then f=

P

​​D
de
|
p
f=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}
; if pnpp_{n}\equiv p and ded_{e} is Pansu differentiable at pp, then f=

P

​​D
de
|
p
f=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}
.

Proof.

A simple computation shows that, if pn,qn𝔾p_{n},q_{n}\in\mathbb{G} and ϵn(0,+)\epsilon_{n}\in(0,+\infty) satisfy (6), then

d(qn,x)d(qn,e)=de(pnδϵnx)de(pn)ϵn.d(q_{n},x)-d(q_{n},e)=\frac{d_{e}(p_{n}\delta_{\epsilon_{n}}x)-d_{e}(p_{n})}{\epsilon_{n}}.

Therefore, if qnfh(𝔾,d)q_{n}\to f\in\partial_{h}(\mathbb{G},d), then we take ϵn:=d(e,qn)1\epsilon_{n}:=d(e,q_{n})^{-1}, which converges to 0, and pn=δϵnqn1Bp_{n}=\delta_{\epsilon_{n}}q_{n}^{-1}\in\partial B. Then (5) holds and, up to passing to a subsequence, pnp_{n} converges to a point pBp\in\partial B.

The opposite direction is also clear. ∎

2.7. Horofunctions on vertical fibers

From the basic ingredients above, we can deduce that all horofunctions are constant on vertical fibers, when a Lipschitz property holds for de:xd(e,x)d_{e}:x\mapsto d(e,x).

Notice that, by [15, Proposition 3.3 and Theorem A.1], the Lipschitz property 7 is satisfied for all homogeneous distances on 𝔾\mathbb{G}, whenever 𝔤\mathfrak{g} is strongly bracket generating, that is, the stratification V1V2V_{1}\oplus V_{2} of 𝔤\mathfrak{g} is such that, for every vV1{0}v\in V_{1}\setminus\{0\}, [v,V1]=V2[v,V_{1}]=V_{2}. The Heisenberg group \mathbb{H} is an example of such groups.

Proposition 2.4 (Vertical invariance of horofunctions).

Suppose that 𝔾\mathbb{G} is a Carnot group and dd a homogeneous distance satisfying

(7) there is L>0 such that |de(x)de(y)|Lρ(x,y) for all x,yB(e,2)B(e,1/2),\text{there is $L>0$ such that }|d_{e}(x)-d_{e}(y)|\leq L\rho(x,y)\text{ for all }x,y\in B(e,2)\setminus B(e,1/2),

for some Riemannian distance ρ\rho on 𝔾\mathbb{G}. Then, horofunctions of (𝔾,d)(\mathbb{G},d) are constant along the cosets of the center [𝔾,𝔾][\mathbb{G},\mathbb{G}]. In particular, for every fh(𝔾,d)f\in\partial_{h}(\mathbb{G},d) there is f^C(𝔾/[𝔾,𝔾])\hat{f}\in C(\mathbb{G}/[\mathbb{G},\mathbb{G}]) such that f=f^πf=\hat{f}\circ\pi.

Proof.

Let ρ\rho be a left-invariant Riemannian metric on 𝔾\mathbb{G}. Recall that, by the Ball-Box Theorem [19, 18, 8, 20], if ζ[𝔾,𝔾]\zeta\in[\mathbb{G},\mathbb{G}] then limϵ0+ρ(e,δϵζ)ϵ=0\lim_{\epsilon\to 0^{+}}\frac{\rho(e,\delta_{\epsilon}\zeta)}{\epsilon}=0. Now, fix fh(𝔾,d)f\in\partial_{h}(\mathbb{G},d), and let pnBp_{n}\in\partial B and ϵn0\epsilon_{n}\to 0 as in Lemma 2.3. Then, for every ζ[𝔾,𝔾]\zeta\in[\mathbb{G},\mathbb{G}] and x𝔾x\in\mathbb{G},

f(xζ)f(x)\displaystyle f(x\zeta)-f(x) =limnde(pnδϵn(xζ))de(pnδϵnx)ϵn\displaystyle=\lim_{n\to\infty}\frac{d_{e}(p_{n}\delta_{\epsilon_{n}}(x\zeta))-d_{e}(p_{n}\delta_{\epsilon_{n}}x)}{\epsilon_{n}}
=limnd((pnδϵnx)1,δϵnζ)d((pnδϵnx)1,e)ϵn\displaystyle=\lim_{n\to\infty}\frac{d((p_{n}\delta_{\epsilon_{n}}x)^{-1},\delta_{\epsilon_{n}}\zeta)-d((p_{n}\delta_{\epsilon_{n}}x)^{-1},e)}{\epsilon_{n}}
Llim supnρ(e,δϵnζ)ϵn=0.\displaystyle\leq L\limsup_{n\to\infty}\frac{\rho(e,\delta_{\epsilon_{n}}\zeta)}{\epsilon_{n}}=0.\qed
Remark 2.5.

We give an example where horofunctions are not constant along the center. Endow the stratified group 𝔾=×\mathbb{G}=\mathbb{H}\times\mathbb{R} with a homogeneous distance of the form

d((0,0,0;0),(x,y,z;t))=|x|+|y|+c|z|+|t|d((0,0,0;0),(x,y,z;t))=|x|+|y|+c\sqrt{|z|}+|t|

with c>0c>0 chosen so that dd satisfies the triangular inequality. Using the notation of the above proof, take

pn\displaystyle p_{n} =(0,0,1/n;1c/n),\displaystyle=(0,0,1/n;1-c/\sqrt{n}), x\displaystyle x =0,\displaystyle=0, ζ\displaystyle\zeta =(0,0,1;0),\displaystyle=(0,0,1;0), ϵn\displaystyle\epsilon_{n} =1/n.\displaystyle=1/\sqrt{n}.

Then, d(e,pn)=1d(e,p_{n})=1 for all nn and

de(pnδϵn(xζ))de(pnδϵnx)ϵn=c(21)0\frac{d_{e}(p_{n}\delta_{\epsilon_{n}}(x\zeta))-d_{e}(p_{n}\delta_{\epsilon_{n}}x)}{\epsilon_{n}}=c(\sqrt{2}-1)\neq 0

for all nn. Finally, a subsequence of qn=δ1/ϵnpnq_{n}=\delta_{1/\epsilon_{n}}p_{n} converges to a horofunction ff which satisfies f(ζ)f(0)0f(\zeta)-f(0)\neq 0, i.e., it is not constant along [𝔾,𝔾][\mathbb{G},\mathbb{G}].

3. Blow-ups of sets and functions in homogeneous groups

As we observed in Lemma 2.3, in homogeneous groups there is a connection between horofunctions in the boundary and directional derivatives along the unit sphere. Wherever the unit sphere is smooth, this directional derivative is the Pansu derivative. While the unit sphere is Pansu differentiable almost everywhere, the nonsmooth points must be studied using a different strategy. In this section, we overview the Kuratowski convergence of closed sets, sometimes credited to Kuratowski–Painlevé, and we use it define the blow-up of functions.

3.1. Kuratowski limits in metric spaces

Let (X,d)(X,d) be a locally compact metric space and let 𝙲𝙻(X)\mathtt{CL}(X) be the family of all closed subsets of XX. If xXx\in X and CXC\subset X, we set d(x,C):=inf{d(x,y):yC}d(x,C):=\inf\{d(x,y):y\in C\}. The Kuratowski limit inferior of a sequence {Cn}n𝙲𝙻(X)\{C_{n}\}_{n\in\mathbb{N}}\subset\mathtt{CL}(X) is defined to be

𝙻𝚒nCn\displaystyle\operatorname*{\mathtt{Li}}_{n\to\infty}C_{n} :={qX:lim supnd(q,Cn)=0}\displaystyle:=\left\{q\in X:\limsup_{n\to\infty}d(q,C_{n})=0\right\}
={qX:nxnCn s.t. limnxn=q},\displaystyle=\left\{q\in X:\forall n\in\mathbb{N}\;\exists x_{n}\in C_{n}\text{ s.t.\leavevmode\nobreak\ }\lim_{n\to\infty}x_{n}=q\right\},

while the Kuratowski limit superior is defined to be

𝙻𝚜nCn\displaystyle\operatorname*{\mathtt{Ls}}_{n\to\infty}C_{n} :={qX:lim infnd(q,Cn)=0}\displaystyle:=\left\{q\in X:\liminf_{n\to\infty}d(q,C_{n})=0\right\}
={qX:N infinite kNxkCk s.t. limkxk=q},\displaystyle=\left\{q\in X:\exists N\subset\mathbb{N}\text{ infinite }\forall k\in N\;\exists x_{k}\in C_{k}\text{ s.t.\leavevmode\nobreak\ }\lim_{k\to\infty}x_{k}=q\right\},

It is clear that 𝙻𝚒nCn𝙻𝚜nCn\operatorname*{\mathtt{Li}}_{n}C_{n}\subseteq\operatorname*{\mathtt{Ls}}_{n}C_{n} and that they are both closed.

If 𝙻𝚒Cn=𝙻𝚜Cn=C\operatorname*{\mathtt{Li}}C_{n}=\operatorname*{\mathtt{Ls}}C_{n}=C, then we say that the CC is the Kuratowski limit of {Cn}n\{C_{n}\}_{n} and we write

C=KlimnCn.C=\operatorname*{K-lim}_{n\to\infty}C_{n}.

If, for all nn\in\mathbb{N}, ΩnX\Omega_{n}\subset X are closed sets and fn:Ωnf_{n}:\Omega_{n}\to\mathbb{R} continuous functions, then we say that, for some ΩX\Omega\subset X closed and f:Ωf:\Omega\to\mathbb{R} continuous,

Klimn(Ωn,fn)=(Ω,f)\operatorname*{K-lim}_{n\to\infty}(\Omega_{n},f_{n})=(\Omega,f)

if Ω=KlimnΩn\Omega=\operatorname*{K-lim}_{n}\Omega_{n} and if, for every xΩx\in\Omega and every sequence {xn}n\{x_{n}\}_{n\in\mathbb{N}} with xnΩnx_{n}\in\Omega_{n} and xnxx_{n}\to x, we have f(x)=limnfn(xn)f(x)=\lim_{n}f_{n}(x_{n}). Notice that this is equivalent to say that

Klimn{(x,fn(x)):xΩn}={(x,f(x)):xΩ}.\operatorname*{K-lim}_{n\to\infty}\{(x,f_{n}(x)):x\in\Omega_{n}\}=\{(x,f(x)):x\in\Omega\}.

If Cn1,,CnJC^{1}_{n},\dots,C^{J}_{n} are sequences of closed sets,

then one easily checks that

𝙻𝚜nj=1JCnjj=1J𝙻𝚜nCnj,andj=1J𝙻𝚒nCnj𝙻𝚒nj=1JCnj.\operatorname*{\mathtt{Ls}}_{n\to\infty}\bigcup_{j=1}^{J}C^{j}_{n}\subset\bigcup_{j=1}^{J}\operatorname*{\mathtt{Ls}}_{n\to\infty}C^{j}_{n},\quad\text{and}\quad\bigcup_{j=1}^{J}\operatorname*{\mathtt{Li}}_{n\to\infty}C^{j}_{n}\subset\operatorname*{\mathtt{Li}}_{n\to\infty}\bigcup_{j=1}^{J}C^{j}_{n}.

Therefore, if the limit KlimnCnj\operatorname*{K-lim}_{n\to\infty}C^{j}_{n} exists for each jj, then we have

(8) Klimnj=1JCnj=j=1JKlimnCnj.\operatorname*{K-lim}_{n\to\infty}\bigcup_{j=1}^{J}C^{j}_{n}=\bigcup_{j=1}^{J}\operatorname*{K-lim}_{n\to\infty}C^{j}_{n}.

It is a classical result of Zarankiewicz that under mild conditions, 𝙲𝙻(X)\mathtt{CL}(X) is sequentially compact with respect to Kuratowski convergence.

Theorem 3.1 (Zarankiewicz [25]).

If (X,d)(X,d) is a separable metric space, then the family of closed sets is sequentially compact with respect to the Kuratowski convergence, that is, if {Cn}n\{C_{n}\}_{n\in\mathbb{N}} is a sequence of closed sets, then there is NN\subset\mathbb{N} infinite and CXC\subset X closed such that KlimNnCn=C\operatorname*{K-lim}_{N\ni n\to\infty}C_{n}=C.

For ϵ0\epsilon\geq 0 and CXC\subset X, let

𝒩ϵ(𝒞):={𝓍:𝒹(𝓍,𝒞)ϵ}, and 𝒩ϵ(𝒞):={𝓍:𝒹(𝓍,𝒳𝒞)>ϵ}.\cal N_{\epsilon}(C):=\{x:d(x,C)\leq\epsilon\},\text{ and }\cal N_{-\epsilon}(C):=\{x:d(x,X\setminus C)>\epsilon\}.

Notice that 𝒩ϵ(𝒞)=𝒳𝒩ϵ(𝒳𝒞)\cal N_{-\epsilon}(C)=X\setminus\cal N_{\epsilon}(X\setminus C).

A set CXC\subset X is a regular closed set if it is the closure of its interior. If CC is a closed set, then XC¯\overline{X\setminus C} is regular closed. If CC is a regular closed set, then

C=ϵ>0𝒩ϵ(𝒞)=ϵ>0𝒩ϵ(𝒞)¯C=\bigcap_{\epsilon>0}\cal N_{\epsilon}(C)=\overline{\bigcup_{\epsilon>0}\cal N_{-\epsilon}(C)}
Lemma 3.2.

Assume XX to be locally compact. Let fn:Xf_{n}:X\to\mathbb{R} be a sequence of continuous functions locally uniformly converging to f:Xf_{\infty}:X\to\mathbb{R}. Then

(9) {f<0}𝙻𝚒n{fn0}𝙻𝚜n{fn0}{f0}.\{f_{\infty}<0\}\subset\operatorname*{\mathtt{Li}}_{n\to\infty}\{f_{n}\leq 0\}\subset\operatorname*{\mathtt{Ls}}_{n\to\infty}\{f_{n}\leq 0\}\subset\{f_{\infty}\leq 0\}.

In particular, if {f<0}¯={f0}\overline{\{f_{\infty}<0\}}=\{f_{\infty}\leq 0\}, then

Klimn{fn0}={f0}.\operatorname*{K-lim}_{n\to\infty}\{f_{n}\leq 0\}=\{f_{\infty}\leq 0\}.
Proof.

For the first inclusion in (9), let pXp\in X with f(p)<ϵ<0f_{\infty}(p)<-\epsilon<0 for some ϵ0\epsilon\leq 0. Then there is r>0r>0 such that B¯(p,r)\bar{B}(p,r) is compact and f(x)<ϵf_{\infty}(x)<-\epsilon for all xB¯(p,r)x\in\bar{B}(p,r). By the uniform convergence on compact sets, there exist NN\in\mathbb{N} such that fn(p)<ϵ/2<0f_{n}(p)<-\epsilon/2<0 for all n>Nn>N. Therefore, p𝙻𝚒n{fn0}p\in\operatorname*{\mathtt{Li}}_{n\to\infty}\{f_{n}\leq 0\}. For the third inclusion in (9), consider a sequence {pn}nX\{p_{n}\}_{n\in\mathbb{N}}\subset X with pnpp_{n}\to p and fn(pn)0f_{n}(p_{n})\leq 0. Then, by the uniform convergence on compact sets, we have limnfn(pn)=f(p)\lim_{n}f_{n}(p_{n})=f(p) and thus f(p)0f(p)\leq 0. The last statement is a direct consequence of the fact that Kuratowski superior and inferior limits are both closed. ∎

A family X\mathscr{F}\subset\mathbb{R}^{X} is strictly monotone if for every pXp\in X there exists γp:[1,1]X\gamma_{p}:[-1,1]\to X continuous with γp(0)=p\gamma_{p}(0)=p such that tf(γp(t))t\mapsto f(\gamma_{p}(t)) is strictly increasing for every ff\in\mathscr{F}.

Lemma 3.3.

If C(X)\mathscr{F}\subset C(X) is strictly monotone and finite,

then

{max<0}¯={max0}.\overline{\{\max\mathscr{F}<0\}}=\{\max\mathscr{F}\leq 0\}.
Proof.

Let p{max0}p\in\{\max\mathscr{F}\leq 0\} with max(p)=0\max\mathscr{F}(p)=0. Let γp:[1,1]X\gamma_{p}:[-1,1]\to X be continuous with γp(0)=p\gamma_{p}(0)=p such that tf(γp(t))t\mapsto f(\gamma_{p}(t)) is strictly increasing for every ff\in\mathscr{F}. It follows that, for every ff\in\mathscr{F} and t<0t<0, we have f(γ(t))<f(γ(0))max(p)=0f(\gamma(t))<f(\gamma(0))\leq\max\mathscr{F}(p)=0. Then pn=γ(1/n)p_{n}=\gamma(-1/n) is a sequence of points converging to pp with max(pn)<0\max\mathscr{F}(p_{n})<0. We conclude that p{max<0}¯p\in\overline{\{\max\mathscr{F}<0\}}. ∎

Lemma 3.4.

Assume that XX is locally compact. For each jj integer between 11 and JJ\in\mathbb{N}, let {fnj}n\{f^{j}_{n}\}_{n\in\mathbb{N}} be a sequence of continuous functions fnj:Xf^{j}_{n}:X\to\mathbb{R} converging uniformly on compact sets to fj:Xf^{j}_{\infty}:X\to\mathbb{R}. Then the sequence of continuous functions gn:=max{fnj}jg_{n}:=\max\{f^{j}_{n}\}_{j} converges uniformly on compact sets to g:=max{fj}jg_{\infty}:=\max\{f^{j}_{\infty}\}_{j}.

Moreover, if {fj}j=1J\{f^{j}_{\infty}\}_{j=1}^{J} is strictly monotone, then

(10) Klimnj=1J{fnj0}=j=1J{fj0}={g0}.\operatorname*{K-lim}_{n\to\infty}\bigcap_{j=1}^{J}\{f^{j}_{n}\leq 0\}=\bigcap_{j=1}^{J}\{f^{j}_{\infty}\leq 0\}=\{g_{\infty}\leq 0\}.
Proof.

We give a proof only for J=2J=2: the general case can then be proved by induction.

So, we assume J=2J=2. Let KXK\Subset X, ϵ>0\epsilon>0, and let (f1f2)(x)=max{f1(x),f2(x)}(f_{1}\vee f_{2})(x)=\max\{f_{1}(x),f_{2}(x)\}. Then there is NN\in\mathbb{N} such that |fnj(x)fj(x)|<ϵ|f^{j}_{n}(x)-f^{j}_{\infty}(x)|<\epsilon for all xKx\in K and jj. We claim that |(fn1fn2)(x)(f1f2)(x)|<3ϵ|(f^{1}_{n}\vee f^{2}_{n})(x)-(f^{1}_{\infty}\vee f^{2}_{\infty})(x)|<3\epsilon for all xKx\in K. To prove the claim we need to check four cases, which by symmetry reduce to the following two: In the first case, (fn1fn2)(x)=fn1(x)(f^{1}_{n}\vee f^{2}_{n})(x)=f^{1}_{n}(x) and (f1f2)(x)=f1(x)(f^{1}_{\infty}\vee f^{2}_{\infty})(x)=f^{1}_{\infty}(x). Then clearly |(fn1fn2)(x)(f1f2)(x)|<ϵ|(f^{1}_{n}\vee f^{2}_{n})(x)-(f^{1}_{\infty}\vee f^{2}_{\infty})(x)|<\epsilon. In the second case, (fn1fn2)(x)=fn1(x)(f^{1}_{n}\vee f^{2}_{n})(x)=f^{1}_{n}(x) and (f1f2)(x)=f2(x)(f^{1}_{\infty}\vee f^{2}_{\infty})(x)=f^{2}_{\infty}(x). Notice that

0\displaystyle 0 f2(x)f1(x)\displaystyle\leq f^{2}_{\infty}(x)-f^{1}_{\infty}(x)
f2(x)fn2(x)+fn2(x)fn1(x)+fn1(x)f1(x)\displaystyle\leq f^{2}_{\infty}(x)-f^{2}_{n}(x)+f^{2}_{n}(x)-f^{1}_{n}(x)+f^{1}_{n}(x)-f^{1}_{\infty}(x)
f2(x)fn2(x)+fn1(x)f1(x)2ϵ.\displaystyle\leq f^{2}_{\infty}(x)-f^{2}_{n}(x)+f^{1}_{n}(x)-f^{1}_{\infty}(x)\leq 2\epsilon.

Therefore,

|(fn1fn2)(x)(f1f2)(x)|=|fn1(x)f2(x)||fn1(x)f1(x)|+|f1(x)f2(x)|3ϵ.|(f^{1}_{n}\vee f^{2}_{n})(x)-(f^{1}_{\infty}\vee f^{2}_{\infty})(x)|=|f^{1}_{n}(x)-f^{2}_{\infty}(x)|\leq|f^{1}_{n}(x)-f^{1}_{\infty}(x)|+|f^{1}_{\infty}(x)-f^{2}_{\infty}(x)|\leq 3\epsilon.

This proves the claim and the first part of the lemma.

For the equalities in (10), notice that {g<0}¯={g0}\overline{\{g_{\infty}<0\}}=\{g_{\infty}\leq 0\} by the strict monotonicity and Lemma 3.3. Thus, we conclude (10) from Lemma 3.2.

3.2. Blow-ups of sets in homogeneous groups

Let 𝔾\mathbb{G} be a homogeneous group with a homogeneous distance dd. If Ω𝔾\Omega\subset\mathbb{G} is closed, {pn}n𝔾\{p_{n}\}_{n\in\mathbb{N}}\subset\mathbb{G} and {ϵn}n(0,+)\{\epsilon_{n}\}_{n\in\mathbb{N}}\subset(0,+\infty) are sequences, we define the blow-up set

𝙱𝚄(Ω,{pn}n,{ϵn}n):=Klimnδ1/ϵn(pn1Ω),\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}):=\operatorname*{K-lim}_{n\to\infty}\delta_{1/\epsilon_{n}}(p_{n}^{-1}\Omega),

if it exists. We sometimes use also the intermediate blow-up sets

𝙱𝚄(Ω,{pn}n,{ϵn}n)\displaystyle\mathtt{BU}^{-}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) :=𝙻𝚒nδ1/ϵn(pn1Ω),\displaystyle:=\operatorname*{\mathtt{Li}}_{n\to\infty}\delta_{1/\epsilon_{n}}(p_{n}^{-1}\Omega),
𝙱𝚄+(Ω,{pn}n,{ϵn}n)\displaystyle\mathtt{BU}^{+}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) :=𝙻𝚜nδ1/ϵn(pn1Ω),\displaystyle:=\operatorname*{\mathtt{Ls}}_{n\to\infty}\delta_{1/\epsilon_{n}}(p_{n}^{-1}\Omega),

which are always well defined and

(11) 𝙱𝚄(Ω,{pn}n,{ϵn}n)𝙱𝚄+(Ω,{pn}n,{ϵn}n).\mathtt{BU}^{-}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})\subset\mathtt{BU}^{+}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}).
Proposition 3.5.

Let Ω𝔾\Omega\subset\mathbb{G} be a nonempty closed set, {pn}n𝔾\{p_{n}\}_{n\in\mathbb{N}}\subset\mathbb{G} and {ϵn}n(0,+)\{\epsilon_{n}\}_{n\in\mathbb{N}}\subset(0,+\infty) sequences with ϵn0\epsilon_{n}\to 0.

  1. (1)

    𝙱𝚄+(Ω,{pn}n,{ϵn}n)\mathtt{BU}^{+}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})\neq\emptyset, if and only if lim infnd(pn,Ω)ϵn<\liminf_{n\to\infty}\frac{d(p_{n},\Omega)}{\epsilon_{n}}<\infty.

  2. (2)

    If 𝙱𝚄(Ω,{pn}n,{ϵn}n)𝔾\mathtt{BU}^{-}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})\neq\mathbb{G}, then lim supnd(pn,𝔾Ω)ϵn<\limsup_{n\to\infty}\frac{d(p_{n},\mathbb{G}\setminus\Omega)}{\epsilon_{n}}<\infty.

In particular, in case pnpp_{n}\to p then we have:

  1. (1’)

    If pΩp\notin\Omega, then 𝙱𝚄(Ω,{pn}n,{ϵn}n)=\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=\emptyset.

  2. (2’)

    If pΩp\in\Omega^{\circ}, then 𝙱𝚄(Ω,{pn}n,{ϵn}n)=𝔾\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=\mathbb{G}.

Proof.

(1) \Rightarrow Let q𝙱𝚄+(Ω,{pn}n,{ϵn}n)q\in\mathtt{BU}^{+}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}). Then there exists NN\subset\mathbb{N} infinite and a sequence {xk}kNΩ\{x_{k}\}_{k\in N}\subset\Omega such that q=limkδ1/ϵk(pk1xk)q=\lim_{k\to\infty}\delta_{1/\epsilon_{k}}(p_{k}^{-1}x_{k}). Therefore,

lim infnd(pn,Ω)ϵnlim infNkd(pk,xk)ϵk=lim infNkd(e,δ1/ϵk(pk1xk))=d(e,q).\liminf_{n\to\infty}\frac{d(p_{n},\Omega)}{\epsilon_{n}}\leq\liminf_{N\ni k\to\infty}\frac{d(p_{k},x_{k})}{\epsilon_{k}}=\liminf_{N\ni k\to\infty}d(e,\delta_{1/\epsilon_{k}}(p_{k}^{-1}x_{k}))=d(e,q).

\Leftarrow Let NN\subset\mathbb{N} infinite and a sequence {xk}kNΩ\{x_{k}\}_{k\in N}\subset\Omega such that limNkd(pn,xk)ϵn<\lim_{N\ni k\to\infty}\frac{d(p_{n},x_{k})}{\epsilon_{n}}<\infty. Since d(pn,xk)ϵn=d(e,δϵk1(pk1xk))\frac{d(p_{n},x_{k})}{\epsilon_{n}}=d(e,\delta_{\epsilon_{k}^{-1}}(p_{k}^{-1}x_{k})), we can assume, up to passing to a subsequence, that the limit limNkδ1/ϵk(pk1xk)\lim_{N\ni k\to\infty}\delta_{1/\epsilon_{k}}(p_{k}^{-1}x_{k}) esists. Thus, 𝙱𝚄+(Ω,{pn}n,{ϵn}n)\mathtt{BU}^{+}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})\neq\emptyset.

(2) Let q𝔾𝙱𝚄(Ω,{pn}n,{ϵn}n)q\in\mathbb{G}\setminus\mathtt{BU}^{-}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) and define xn:=pnδϵnqx_{n}:=p_{n}\delta_{\epsilon_{n}}q. Since q𝙱𝚄(Ω,{pn}n,{ϵn}n)q\notin\mathtt{BU}^{-}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}), there is NN\subset\mathbb{N} infinite such that xkΩx_{k}\notin\Omega for all kNk\in N. Therefore,

lim supnd(pn,𝔾Ω)ϵnlim supnd(pn,pnδϵnq)ϵn=d(e,q).\limsup_{n\to\infty}\frac{d(p_{n},\mathbb{G}\setminus\Omega)}{\epsilon_{n}}\leq\limsup_{n\to\infty}\frac{d(p_{n},p_{n}\delta_{\epsilon_{n}}q)}{\epsilon_{n}}=d(e,q).

Proposition 3.6.

Let Ω𝔾\Omega\subset\mathbb{G} be a nonempty closed set and pΩp\in\partial\Omega.

Suppose that there exists a neighborhood UU of pp and a finite family of continuous functions Fj:UF_{j}:U\to\mathbb{R} with jJj\in J finite such that ΩU=jJ{Fj0}\Omega\cap U=\bigcap_{j\in J}\{F_{j}\leq 0\} and Fj(p)=0F_{j}(p)=0 for all jj. Suppose also that each FjF_{j} is strictly Pansu differentiable at pp and that

(12) 0𝚌𝚟𝚡{

P

​​D
Fj
|
p
}
jJ
.
0\notin\mathtt{cvx}\{\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}\}_{j\in J}.

Let pnpp_{n}\to p and ϵn0+\epsilon_{n}\to 0^{+}, and assume that 𝙱𝚄(Ω,{pn}n,{ϵn}n)\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) exists. Then

𝙱𝚄(Ω,{pn}n,{ϵn}n)={x𝔾:

P

​​D
Fj
|
p
(x)
tj
,jJ
}
\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=\{x\in\mathbb{G}:\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(x)\leq t_{j},\ j\in J\}

with tj{,+}t_{j}\in\mathbb{R}\cup\{-\infty,+\infty\} defined as follows:

  1. (1)

    if limnd(pn,{Fj0})ϵn=+\lim_{n}\frac{d(p_{n},\{F_{j}\leq 0\})}{\epsilon_{n}}=+\infty, then tj=t_{j}=-\infty;

  2. (2)

    if limnd(pn,𝔾{Fj0})ϵn=+\lim_{n}\frac{d(p_{n},\mathbb{G}\setminus\{F_{j}\leq 0\})}{\epsilon_{n}}=+\infty, then tj=+t_{j}=+\infty;

  3. (3)

    otherwise, there are qnj{Fj=0}q_{n}^{j}\in\{F_{j}=0\} such that, up to a subsequence, limnδ1/ϵn((qnj)1pn)=:vj\lim_{n}\delta_{1/\epsilon_{n}}((q_{n}^{j})^{-1}p_{n})=:v_{j}, and we set tj=

    P

    ​​D
    Fj
    |
    p
    (vj)
    t_{j}=-\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(v_{j})
    .

Proof.

Let pnpp_{n}\to p and ϵn0+\epsilon_{n}\to 0^{+}, assume that 𝙱𝚄(Ω,{pn}n,{ϵn}n)\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) exists.

If there is any jj such that limnd(pn,{Fj0})ϵn=+\lim_{n}\frac{d(p_{n},\{F_{j}\leq 0\})}{\epsilon_{n}}=+\infty then 𝙱𝚄(Ω,{pn}n,{ϵn}n)=\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=\emptyset by Proposition 3.5.

Again by Proposition 3.5, for any jj such that limnd(pn,𝔾{Fj0})ϵn=+\lim_{n}\frac{d(p_{n},\mathbb{G}\setminus\{F_{j}\leq 0\})}{\epsilon_{n}}=+\infty, we know 𝙱𝚄({Fj0},{pn}n,{ϵn}n)=𝔾={

P

​​D
Fj
|
p
+
}
\mathtt{BU}(\{F_{j}\leq 0\},\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=\mathbb{G}=\{\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}\leq+\infty\}
.

Let J^\hat{J} be the set of indices which do not fall into the first two cases. For all jJ^j\in\hat{J}, there are qnj{Fj=0}q_{n}^{j}\in\{F_{j}=0\} with d(pn,qnj)=d(pn,{Fj=0})d(p_{n},q_{n}^{j})=d(p_{n},\{F_{j}=0\}) and lim supnd(pn,qnj)ϵn<\limsup_{n}\frac{d(p_{n},q_{n}^{j})}{\epsilon_{n}}<\infty. Up to a subsequence, we can assume that the limit limnδ1/ϵn(qn1pn)=vj\lim_{n}\delta_{1/\epsilon_{n}}(q_{n}^{-1}p_{n})=v_{j} exists. Define

fnj(x)\displaystyle f_{n}^{j}(x) :=Fj(pnδϵnx)ϵn,\displaystyle:=\frac{F_{j}(p_{n}\delta_{\epsilon_{n}}x)}{\epsilon_{n}},

and note that near δ1/ϵn(pn1p)\delta_{1/\epsilon_{n}}(p_{n}^{-1}p), the locus j{fnj0}\bigcap\nolimits_{j}\{f_{n}^{j}\leq 0\} is a local description of the translated and dilated set δ1/ϵn(pn1Ω)\delta_{1/\epsilon_{n}}(p_{n}^{-1}\Omega) for all n>0n>0. We then observe that

fnj(x)=Fj(pnδϵnx)F(pn)ϵn+F(pn)F(qn)ϵn.f_{n}^{j}(x)=\frac{F_{j}(p_{n}\delta_{\epsilon_{n}}x)-F(p_{n})}{\epsilon_{n}}+\frac{F(p_{n})-F(q_{n})}{\epsilon_{n}}.

By the strict Pansu differentiability of FjF_{j} at pp, the functions fnjf_{n}^{j} converge uniformly on compact sets to fj(x):=

P

​​D
Fj
|
p
(x)
+

P

​​D
Fj
|
p
(vj)
f_{\infty}^{j}(x):=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(x)+\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(v_{j})
.

Condition (12) implies that there is wV1w\in V_{1} such that

P

​​D
Fj
|
p
(w)
>0
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(w)>0
for all jj. Define γ(t)=pexp(tw)\gamma(t)=p\exp(tw). Then

ddtfj(γ(t))=

P

​​D
Fj
|
p
(w)
,
\frac{\,\mathrm{d}}{\,\mathrm{d}t}f_{\infty}^{j}(\gamma(t))=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(w),

which is strictly positive for tt in a neighborhood of 0, for all jJ^j\in\hat{J}. Therefore, the family of functions {fj}jJ^\{f_{\infty}^{j}\}_{j\in\hat{J}} is strictly monotone and we conclude by Lemma 3.4 that

𝙱𝚄(Ω,{pn}n,{ϵn}n)=jJ^{fj0}.\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=\bigcap_{j\in\hat{J}}\{f_{\infty}^{j}\leq 0\}.

Proposition 3.7.

Let Ω𝔾\Omega\subset\mathbb{G} be a nonempty closed set and pΩp\in\partial\Omega. Suppose that there exists a neighborhood UU of pp and a finite family of continuous functions Fj:UF_{j}:U\to\mathbb{R} with jJj\in J such that ΩU=jJ{Fj0}\Omega\cap U=\bigcap_{j\in J}\{F_{j}\leq 0\} and Fj(p)=0F_{j}(p)=0. Suppose also that each FjF_{j} is smooth and that {

P

​​D
Fj
|
p
}
jJ
\{\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}\}_{j\in J}
are linearly independent.

Then, for every (tj)j({+})J(t_{j})_{j}\in(\mathbb{R}\cup\{+\infty\})^{J} and every ϵn0+\epsilon_{n}\to 0^{+}, there are pnpp_{n}\to p such that

(13) 𝙱𝚄(Ω,{pn}n,{ϵn}n)={x𝔾:

P

​​D
Fj
|
p
(x)
tj
,jJ
}
.
\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=\{x\in\mathbb{G}:\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(x)\leq t_{j},\ j\in J\}.
Proof.

Define the function

R(ϵ)=maxjJsup{|Fj(pδηw)Fj(p)η

P

​​D
Fj|p(w)|:wB(0,1), 0<ηϵ}
.
R(\epsilon)=\max_{j\in J}\sup\left\{\left|\frac{F_{j}(p\delta_{\eta}w)-F_{j}(p)}{\eta}-\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(w)\right|:w\in B(0,1),\ 0<\eta\leq\epsilon\right\}.

Since JJ is finite and FjF_{j} are all smooth, we have R(ϵ)=O(ϵ)R(\epsilon)=O(\epsilon), see [7, Theorem 1.42].

Fix (tj)j({+})J(t_{j})_{j}\in(\mathbb{R}\cup\{+\infty\})^{J} and ϵn0+\epsilon_{n}\to 0^{+}. Since {

P

​​D
Fj
|
p
}
jJ
\{\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}\}_{j\in J}
are linearly independent, there are w1,w2𝔾w_{1},w_{2}\in\mathbb{G} such that

P

​​D
Fj
|
p
(w1)
\displaystyle\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(w_{1})
=tj if tj<+,\displaystyle=-t_{j}\text{ if }t_{j}<+\infty,

P

​​D
Fj
|
p
(w1)
\displaystyle\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(w_{1})
=1 if tj=+;\displaystyle=1\text{ if }t_{j}=+\infty;

P

​​D
Fj
|
p
(w2)
\displaystyle\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(w_{2})
=0 if tj<+,\displaystyle=0\text{ if }t_{j}<+\infty,

P

​​D
Fj
|
p
(w2)
\displaystyle\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(w_{2})
=1 if tj=+.\displaystyle=-1\text{ if }t_{j}=+\infty.

Now, we define pn:=p(δϵn3/4w2)(δϵnw1)p_{n}:=p(\delta_{\epsilon_{n}^{3/4}}w_{2})(\delta_{\epsilon_{n}}w_{1}). As in the proof of Proposition 3.6, we define fnj(x)=Fj(pnδϵnx)ϵnf_{n}^{j}(x)=\frac{F_{j}(p_{n}\delta_{\epsilon_{n}}x)}{\epsilon_{n}} and recall that j{fnj0}\bigcap_{j}\{f_{n}^{j}\leq 0\} gives a local description of δ1/ϵn(pn1Ω)\delta_{1/\epsilon_{n}}(p_{n}^{-1}\Omega). In the limit, our choice of pnp_{n} will allow us to express 𝙱𝚄(Ω,{pn}n,{ϵn}n)=j{fj0}\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=\bigcap_{j}\{f^{j}_{\infty}\leq 0\} as in equation (13). Indeed, by our choice of pnp_{n}, for any x𝔾x\in\mathbb{G}, it follows that

fnj(x)=Fj(pnδϵnx)ϵn\displaystyle f_{n}^{j}(x)=\frac{F_{j}(p_{n}\delta_{\epsilon_{n}}x)}{\epsilon_{n}} =Fj(p(δϵn3/4w2)(δϵn(w1x)))Fj(p)ϵn\displaystyle=\frac{F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2})(\delta_{\epsilon_{n}}(w_{1}x)))-F_{j}(p)}{\epsilon_{n}}
=Fj(p(δϵn3/4w2)(δϵn(w1x)))Fj(p(δϵn3/4w2))ϵn+Fj(p(δϵn3/4w2))Fj(p)ϵn,\displaystyle=\frac{F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2})(\delta_{\epsilon_{n}}(w_{1}x)))-F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2}))}{\epsilon_{n}}+\frac{F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2}))-F_{j}(p)}{\epsilon_{n}},

where

limnFj(p(δϵn3/4w2)(δϵn(w1x)))Fj(p(δϵn3/4w2))ϵn=

P

​​D
Fj
|
p
(w1x)
={

P

​​D
Fj
|
p
(x)
tj
 if tj<+
1 if tj=+
\lim_{n\to\infty}\frac{F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2})(\delta_{\epsilon_{n}}(w_{1}x)))-F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2}))}{\epsilon_{n}}=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(w_{1}x)=\begin{cases}\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(x)-t_{j}&\text{ if }t_{j}<+\infty\\ 1&\text{ if }t_{j}=+\infty\end{cases}

and, if tj<t_{j}<\infty,

limn|Fj(p(δϵn3/4w2))Fj(p)ϵn|=limn|Fj(p(δϵn3/4w2))Fj(p)ϵn3/4|1ϵn1/4limnR(ϵn3/4)ϵn1/4=0,\lim_{n\to\infty}\left|\frac{F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2}))-F_{j}(p)}{\epsilon_{n}}\right|=\lim_{n\to\infty}\left|\frac{F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2}))-F_{j}(p)}{\epsilon_{n}^{3/4}}\right|\frac{1}{\epsilon_{n}^{1/4}}\\ \leq\lim_{n\to\infty}\frac{R(\epsilon_{n}^{3/4})}{\epsilon_{n}^{1/4}}=0,

while, if tj=t_{j}=\infty, then

limnFj(p(δϵn3/4w2))Fj(p)ϵn=limnFj(p(δϵn3/4w2))Fj(p)ϵn3/41ϵn1/4=

P

​​D
Fj
|
p
(w2)limn1ϵn1/4
=
.
\lim_{n\to\infty}\frac{F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2}))-F_{j}(p)}{\epsilon_{n}}=\lim_{n\to\infty}\frac{F_{j}(p(\delta_{\epsilon_{n}^{3/4}}w_{2}))-F_{j}(p)}{\epsilon_{n}^{3/4}}\frac{1}{\epsilon_{n}^{1/4}}\\ =\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{j}|_{p}(w_{2})\lim_{n\to\infty}\frac{1}{\epsilon_{n}^{1/4}}=-\infty.

Finally, using the same strategy as in the second part of the proof of Proposition 3.6, we conclude that (13) holds. ∎

3.3. Blow-ups of functions in homogeneous groups

For a continuous function f:Ωf:\Omega\to\mathbb{R}, we define

𝙱𝚄((Ω,f),{pn}n,{ϵn}n):=Klimn(δ1/ϵn(pn1Ω),f(pnδϵn)f(pn)ϵn).\mathtt{BU}((\Omega,f),\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}):=\operatorname*{K-lim}_{n\to\infty}\left(\delta_{1/\epsilon_{n}}(p_{n}^{-1}\Omega),\frac{f(p_{n}\delta_{\epsilon_{n}}\cdot)-f(p_{n})}{\epsilon_{n}}\right).
Proposition 3.8.

Let Ω𝔾\Omega\subset\mathbb{G} be a nonempty closed set, {pn}n𝔾\{p_{n}\}_{n\in\mathbb{N}}\subset\mathbb{G} and {ϵn}n(0,+)\{\epsilon_{n}\}_{n\in\mathbb{N}}\subset(0,+\infty) sequences with pnpΩp_{n}\to p\in\Omega and ϵn0\epsilon_{n}\to 0. Suppose that Ω0:=𝙱𝚄(Ω,{pn}n,{ϵn}n)\Omega_{0}:=\mathtt{BU}(\Omega,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) exists. Let f:𝔾f:\mathbb{G}\to\mathbb{R} be a continuous function that is strictly Pansu differentiable at pp. Then

𝙱𝚄((Ω,f),{pn}n,{ϵn}n)=(Ω0,

P

​​D
f(p)
|
Ω0
)
.
\mathtt{BU}((\Omega,f),\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=(\Omega_{0},\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}f(p)|_{\Omega_{0}}).
Proof.

Let fn(x):=f(pnδϵnx)f(pn)ϵnf_{n}(x):=\frac{f(p_{n}\delta_{\epsilon_{n}}x)-f(p_{n})}{\epsilon_{n}}. If xnδ1/ϵn(pn1Ω)x_{n}\in\delta_{1/\epsilon_{n}}(p_{n}^{-1}\Omega) are such that xnxΩ0x_{n}\to x\in\Omega_{0}, then fn(xn)

P

​​D
f
|
p
[x]
f_{n}(x_{n})\to\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}f|_{p}[x]
, by the strict Pansu differentiability of ff at pp. ∎

If QQ is a closed set, we say that a function f:Qf:Q\to\mathbb{R} is smooth if there exists a smooth extension of ff in a neighborhood of QQ. In particular, the derivative of ff at points pQp\in\partial Q is well defined.

Theorem 3.9.

Let Ω𝔾\Omega\subset\mathbb{G} be a closed set such that there is a family 𝒬\cal Q of regular closed sets with disjoint interiors such that Ω=Q𝒬Q\Omega=\bigcup_{Q\in\cal Q}Q. For each Q𝒬Q\in\cal Q, let fQ:𝔾f_{Q}:\mathbb{G}\to\mathbb{R} smooth such that the function f:Ωf:\Omega\to\mathbb{R} defined by

f(x):=χ(x)Q𝒬fQ(x)𝟙Q(x)f(x):=\chi(x)\sum_{Q\in\cal Q}f_{Q}(x){\mathds{1}\!}_{Q}(x)

is Lipschitz continuous, where χ(x):=(Q𝒬𝟙Q(x))1\chi(x):=\left(\sum_{Q\in\cal Q}{\mathds{1}\!}_{Q}(x)\right)^{-1}.

Let {pn}n𝔾\{p_{n}\}_{n\in\mathbb{N}}\subset\mathbb{G} and {ϵn}n(0,+)\{\epsilon_{n}\}_{n\in\mathbb{N}}\subset(0,+\infty) sequences with pnpΩp_{n}\to p\in\Omega^{\circ} and ϵn0\epsilon_{n}\to 0. Assume that RQ:=𝙱𝚄(Q,{pn}n,{ϵn}n)R_{Q}:=\mathtt{BU}(Q,\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) exists for every Q𝒬Q\in\cal Q.

Then

𝔾=Q𝒬RQ\mathbb{G}=\bigcup_{Q\in\cal Q}R_{Q}

and 𝙱𝚄((Ω,f),{pn}n,{ϵn}n)=(𝔾,g)\mathtt{BU}((\Omega,f),\{p_{n}\}_{n},\{\epsilon_{n}\}_{n})=(\mathbb{G},g) exists, where

(14) g(x)=χ~(x)Q𝒬(

P

​​D
fQ
|
p
(x)
+cQ
)
𝟙RQ(x)
,
g(x)=\tilde{\chi}(x)\sum_{Q\in\cal Q}\left(\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}f_{Q}|_{p}(x)+c_{Q}\right){\mathds{1}\!}_{R_{Q}}(x),

with χ~(x):=(Q𝒬𝟙RQ(x))1\tilde{\chi}(x):=\left(\sum_{Q\in\cal Q}{\mathds{1}\!}_{R_{Q}}(x)\right)^{-1} and cQc_{Q}\in\mathbb{R}.

Notice that the constants cQc_{Q} can be determined by the continuity of gg and g(e)=0g(e)=0. If there are more than one choice of such constants, the resulting function is still the same: indeed, if gg and gg^{\prime} are two functions as in (14) with different constants, then ggg-g^{\prime} is a piecewise constant and continuous function that is 0 in ee, and thus g=gg=g^{\prime}. Moreover, we remark that we don’t need the limit sets RQR_{Q} to have disjoint interiors.

Proof.

The fact that 𝔾=Q𝒬RQ\mathbb{G}=\bigcup_{Q\in\cal Q}R_{Q} follows from pΩp\in\Omega^{\circ} and (8). Next, set gn:δ1/ϵn(pn1Ω)g_{n}:\delta_{1/\epsilon_{n}}(p_{n}^{-1}\Omega)\to\mathbb{R}, gn(x):=f(pnδϵnx)f(pn)ϵng_{n}(x):=\frac{f(p_{n}\delta_{\epsilon_{n}}x)-f(p_{n})}{\epsilon_{n}}. The family of functions {gn}n\{g_{n}\}_{n\in\mathbb{N}} is uniformly Lipschitz and gn(e)=0g_{n}(e)=0 for all nn. Thus, the set

𝒩:={𝒩 infinite:{𝓃}𝓃𝒩 converge}\cal N:=\{N\subset\mathbb{N}\text{ infinite}:\{g_{n}\}_{n\in N}\text{ converge}\}

is nonempty and for every NN\subset\mathbb{N} infinite there is N𝒩N^{\prime}\in\cal N with NNN^{\prime}\subset N. For every N𝒩N\in\cal N, define gN:=limNngng^{N}:=\lim_{N\ni n\to\infty}g_{n}. We aim to prove that gN=gg^{N}=g for all N𝒩N\in\cal N.

Let xRQx\in R_{Q} for some Q𝒬Q\in\cal Q. Then there exist ynQy_{n}\in Q such that xn:=δ1/ϵn(pn1yn)xx_{n}:=\delta_{1/\epsilon_{n}}(p_{n}^{-1}y_{n})\to x. Therefore, gn(xn)g(x)g_{n}(x_{n})\to g(x), where

gn(xn)\displaystyle g_{n}(x_{n}) =f(yn)f(pn)ϵn\displaystyle=\frac{f(y_{n})-f(p_{n})}{\epsilon_{n}}
=fQ(yn)fQ(pn)ϵn+fQ(pn)f(pn)ϵn.\displaystyle=\frac{f_{Q}(y_{n})-f_{Q}(p_{n})}{\epsilon_{n}}+\frac{f_{Q}(p_{n})-f(p_{n})}{\epsilon_{n}}.

Since fQf_{Q} is smooth at pp, we have limnfQ(yn)fQ(pn)ϵn=

P

​​D
fQ
|
p
[x]
\lim_{n}\frac{f_{Q}(y_{n})-f_{Q}(p_{n})}{\epsilon_{n}}=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}f_{Q}|_{p}[x]
. Therefore, if N𝒩N\in\cal N, then the limit cQN:=limnfQ(pn)f(pn)ϵnc_{Q}^{N}:=\lim_{n}\frac{f_{Q}(p_{n})-f(p_{n})}{\epsilon_{n}} exists and it is equal to gN(x)

P

​​D
fQ
|
p
[x]
g^{N}(x)-\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}f_{Q}|_{p}[x]
. Moreover,

gN(x)=χ~(x)Q𝒬(

P

​​D
fQ
|
p
(x)
+cQN
)
𝟙RQ(x)
.
g^{N}(x)=\tilde{\chi}(x)\sum_{Q\in\cal Q}\left(\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}f_{Q}|_{p}(x)+c^{N}_{Q}\right){\mathds{1}\!}_{R_{Q}}(x).

Finally, gNg^{N} is continuous and gN(e)=0g^{N}(e)=0.

So, for any pair N,N𝒩N,N^{\prime}\in\cal N, the difference gNgNg^{N}-g^{N^{\prime}} is a piecewise constant and continuous function that takes the value 0 at ee. Hence, gNgN0g^{N}-g^{N^{\prime}}\equiv 0, for all N𝒩N\in\cal N.

This theorem will allow us to finish our description of the horofunction boundary. At non-smooth points, horofunctions do not necessarily correspond to Pansu derivatives, but instead are piecewise defined by Pansu derivatives in each blow-up region. Theorem 3.9 can also be used to recover results about the horofunction boundaries of normed spaces as in [9, 23].

4. Vertical sequences in the Heisenberg group \mathbb{H}

In this section, we focus on the Heisenberg group, see Section 2.4. We extend to sub-Finsler distances a result that Klein–Nicas proved for the sub-Riemannian and the Korany distances in [13, 12]. In particular, we show that, for any sub-Finsler metric in the Heisenberg group \mathbb{H}, vertical sequences induce a topological disk in the horoboundary. The result is not true for all homogeneous distances in \mathbb{H}, see Remark 4.4.

Theorem 4.1.

Let dd be the sub-Finsler distance on \mathbb{H} generated by norm \|\cdot\| on the horizontal plane. Let {wn}n2\{w_{n}\}_{n\in\mathbb{N}}\subset\mathbb{R}^{2} be a bounded sequence and {sn}n\{s_{n}\}_{n\in\mathbb{N}}\subset\mathbb{R} with |sn||s_{n}|\to\infty, and set pn=(wn,sn)p_{n}=(w_{n},s_{n})\in\mathbb{H}. Then for all (v,t)(v,t)\in\mathbb{H}

limnd(pn,(v,t))d(pn,e)(wnwnv)=0.\lim_{n\to\infty}d(p_{n},(v,t))-d(p_{n},e)-(\|w_{n}\|-\|w_{n}-v\|)=0.

There is, therefore, a topological disk {pwπ(p)w:w2}C()\{p\mapsto\|w\|-\|\pi(p)-w\|:w\in\mathbb{R}^{2}\}\subset C(\mathbb{H}) in the horofunction boundary.

We need a couple of lemmas before the proof of the theorem. We start with a technical lemma concerning convex geometry. Fix b>0b>0 and an open bounded convex set Q2Q\subset\mathbb{R}^{2}. Dilate QQ by λ1\lambda\geq 1, and take two points p,q(λQ)p,q\in\partial(\lambda Q) so that |pq|b|p-q|\leq b. The line passing through pp and qq cuts λQ\lambda Q into two parts with areas ss and tt respectively, say sts\leq t. Then the lemma says there is MM such that that s<Ms<M for all λ1\lambda\geq 1.

Lemma 4.2.

Let Q2Q\subset\mathbb{R}^{2} be an open bounded convex set. Fix b>0b>0 and define for λ1\lambda\geq 1

x¯(λ)\displaystyle\bar{x}(\lambda) :=inf{x:L1{y:(x,y)λQ}b},\displaystyle:=\inf\{x\in\mathbb{R}:L^{1}\{y:(x,y)\in\lambda Q\}\geq b\},
Qλ\displaystyle Q^{-}_{\lambda} :={(x,y)λQ:xx¯(λ)}.\displaystyle:=\{(x,y)\in\lambda Q:x\leq\bar{x}(\lambda)\}.

Then

sup{L2(Qλ):λ1}<,\sup\{L^{2}(Q^{-}_{\lambda}):\lambda\geq 1\}<\infty,

where L1L^{1} and L2L^{2} denote the 1- and 2-dimensional Lebesgue measures, respectively.

Proof.

Since L2(Qλ)λ2L2(Q)L^{2}(Q^{-}_{\lambda})\leq\lambda^{2}L^{2}(Q), we must show that L2(Qλ)L^{2}(Q^{-}_{\lambda}) remains bounded for λ\lambda large. Taking λ\lambda large enough, we can assume x¯(λ)<+\bar{x}(\lambda)<+\infty. Define

Vλ(x)=L1{y:(x,y)λQ},V_{\lambda}(x)=L^{1}\{y:(x,y)\in\lambda Q\},

and note that Vλ(x)=λV1(x/λ)V_{\lambda}(x)=\lambda V_{1}(x/\lambda). Up to translating QQ, we can assume V1(x)=0V_{1}(x)=0 for all x0x\leq 0 and V1(x)>0V_{1}(x)>0 for small x>0x>0. Moreover, since QQ is convex, V1V_{1} is a concave function. If limx0+V1(x)>0\lim_{x\to 0^{+}}V_{1}(x)>0, then x¯(λ)=0\bar{x}(\lambda)=0 and thus L2(Qλ)=0L^{2}(Q^{-}_{\lambda})=0 for λ\lambda large.

Now assuming that limx0+V1(x)=0\lim_{x\to 0^{+}}V_{1}(x)=0, we have that x¯(λ)>0\bar{x}(\lambda)>0 for all λ\lambda. By concavity, there are ϵ,m>0\epsilon,m>0 such that V1(x)mxV_{1}(x)\geq mx for all x(0,ϵ]x\in(0,\epsilon]. By the definition of x¯\bar{x}, if x<x¯(λ)x<\bar{x}(\lambda) then Vλ(x)bV_{\lambda}(x)\leq b. For λ\lambda large, Vλ(λϵ)=λV1(ϵ)λmϵbV_{\lambda}(\lambda\epsilon)=\lambda V_{1}(\epsilon)\geq\lambda m\epsilon\geq b, and so x¯(λ)<λϵ\bar{x}(\lambda)<\lambda\epsilon. It follows that

b>Vλ(x¯(λ)/2)mx¯(λ)/2,b>V_{\lambda}(\bar{x}(\lambda)/2)\geq m\bar{x}(\lambda)/2,

that is, x¯(λ)2b/m\bar{x}(\lambda)\leq 2b/m. We conclude that

L2(Cλ)=0x¯(λ)Vλ(x)dx2b2m,L^{2}(C^{-}_{\lambda})=\int_{0}^{\bar{x}(\lambda)}V_{\lambda}(x)\,\mathrm{d}x\leq\frac{2b^{2}}{m},

for λ\lambda large enough. ∎

Lemma 4.3.

For any sub-Finsler metric dd on \mathbb{H} and any v2v\in\mathbb{R}^{2},

(15) limt+[de((v,t))de((0,t))+de((v,0))]=0.\lim_{t\to+\infty}\left[d_{e}((v,t))-d_{e}((0,t))+d_{e}((v,0))\right]=0.

Moreover, the convergence is uniform in vv on compact sets.

Proof.

By the triangle inequality, we have

(16) de((v,t))de((0,t))+de((v,0))0d_{e}((v,t))-d_{e}((0,t))+d_{e}((v,0))\geq 0

for all tt. Let Q2Q^{*}\subset\mathbb{R}^{2} be the convex set dual to the unit ball QQ of the norm \lVert\cdot\rVert on 2\mathbb{R}^{2}. Let II be the rotation by π2\frac{\pi}{2} of QQ^{*}.

Define a=a(t)=de((v,t))a=a(t)=d_{e}((v,t)), b=de((v,0))b=d_{e}((v,0)) and h=h(t)=de((0,t))h=h(t)=d_{e}((0,t)). For tt large enough, the projection γ1:[0,1]2\gamma_{1}:[0,1]\to\mathbb{R}^{2} of a geodesic from (0,0)(0,0) to (v,t)(v,t) is a portion of the boundary of λI\lambda I, for some λ\lambda, with γ1(0)=(0,0)\gamma_{1}(0)=(0,0) and γ1(1)=v\gamma_{1}(1)=v. Notice that aa is the length of γ1\gamma_{1}, that b=vb=\lVert v\rVert is the length of a chord of (λI)\partial(\lambda I) and that tt is the area one of the two parts of λI\lambda I separated by the line passing through 0 and vv. Let ss be the area of the other part and cc the length of (λI)γ1\partial(\lambda I)\setminus\gamma_{1}. If AA is the area of II and \ell is the length of I\partial I, we have a+c=λa+c=\lambda\ell and t+s=λ2A.t+s=\lambda^{2}A. See Figure 2.

γ1\gamma_{1}bbvv0ssttttγ2\gamma_{2}
Figure 2. Convex geometry and vertical sequences

The projection γ2\gamma_{2} of a geodesic from (0,0)(0,0) to (0,t)(0,t) is the boundary of μC\mu C, for some μ\mu so that t=L2(μC)=μ2At=L^{2}(\mu C)=\mu^{2}A. Then h=de((0,t))h=d_{e}((0,t)) is the length of the boundary of μC\mu C. Therefore,

(17) h=μλ(a+c)=tt+s(a+c)tt+s(a+b).h=\frac{\mu}{\lambda}(a+c)=\sqrt{\frac{t}{t+s}}(a+c)\geq\sqrt{\frac{t}{t+s}}(a+b).

By Lemma 4.2, there is M>0M>0 such that s<Ms<M for all tt sufficiently large. Thus, by combining (16) and (17), we see that h(t)h(t) converges to a(t)+ba(t)+b, completing the first part of the proof.

For the uniform convergence, if we define ft(v)=de((v,t))de((0,t))+de((v,0))f_{t}(v)=d_{e}((v,t))-d_{e}((0,t))+d_{e}((v,0)), then by the reverse triangle inequality, ft:(2×{0},d)f_{t}:(\mathbb{R}^{2}\times\{0\},d)\to\mathbb{R} is Lipschitz, i.e.,

|ft(v)ft(w)|d((v,t),(w,t))+d((v,0),(w,0))=2d((v,0),(w,0)),|f_{t}(v)-f_{t}(w)|\leq d((v,t),(w,t))+d((v,0),(w,0))=2d((v,0),(w,0)),

and ft(e)=0f_{t}(e)=0. Therefore, the pointwise convergence is uniform on compact sets. ∎

Proof of Theorem 4.1.

It suffices to consider the case when sn+s_{n}\to+\infty. Notice that

d(pn,(v,t))d(pn,e)(wnwnv)=de((wnv,sntω(v,wn)/2))de((0,sntω(v,wn)/2))+de((wnv,0))de((wn,sn))+de((0,sn))de((wn,0))+de((0,sntω(v,wn)/2))de((0,sn)),d(p_{n},(v,t))-d(p_{n},e)-(\|w_{n}\|-\|w_{n}-v\|)\\ =d_{e}((w_{n}-v,s_{n}-t-\omega(v,w_{n})/2))-d_{e}((0,s_{n}-t-\omega(v,w_{n})/2))+d_{e}((w_{n}-v,0))\\ -d_{e}((w_{n},s_{n}))+d_{e}((0,s_{n}))-d_{e}((w_{n},0))\\ +d_{e}((0,s_{n}-t-\omega(v,w_{n})/2))-d_{e}((0,s_{n})),

where, ω\omega is the standard symplectic form on 2\mathbb{R}^{2}. Using Lemma 4.3 and the boundedness of wnw_{n},

limnde((wnv,sntω(v,wn)/2))de((0,sntω(v,wn)/2))+de((wnv,0))=0,\lim_{n\to\infty}d_{e}((w_{n}-v,s_{n}-t-\omega(v,w_{n})/2))-d_{e}((0,s_{n}-t-\omega(v,w_{n})/2))+d_{e}((w_{n}-v,0))=0,

and

limnde((wn,sn))+de((0,sn))de((wn,0))=0.\lim_{n\to\infty}-d_{e}((w_{n},s_{n}))+d_{e}((0,s_{n}))-d_{e}((w_{n},0))=0.

Finally,

limnde((0,sntω(v,wn)/2))de((0,sn))=de((0,1))limn(sntω(v,wn)/2sn)=0.\lim_{n\to\infty}d_{e}((0,s_{n}-t-\omega(v,w_{n})/2))-d_{e}((0,s_{n}))=d_{e}((0,1))\lim_{n\to\infty}\left(\sqrt{s_{n}-t-\omega(v,w_{n})/2}-\sqrt{s_{n}}\right)=0.

For the last statement, fix w2w\in\mathbb{R}^{2} and set pn=(w,n)p_{n}=(w,n)\in\mathbb{H}. Then pnf(v,t)=wwvp_{n}\to f(v,t)=\lVert w\rVert-\lVert w-v\rVert in the horofunction boundary. ∎

Remark 4.4.

For general homogeneous distances, Lemma 4.3 is not true. As an example, consider the function f:f:\mathbb{R}\to\mathbb{R} defined by

f(x):=(1)k(x)(2|x|31+k(x)), where k(x):=log(x)log(3),f(x):=(-1)^{k(x)}\left(2|x|-3^{1+k(x)}\right),\text{ where }k(x):=\left\lfloor\frac{\log(x)}{\log(3)}\right\rfloor,

which is piecewise linear with derivative ±2\pm 2 and satisfies |x|f(x)|x|-|x|\leq f(x)\leq|x| for all xx.

[Uncaptioned image]

Consider the function ϕ(v):=f(|v|)\phi(v):=f(|v|) on the disk in 2\mathbb{R}^{2}. Since ϕ\phi is Lipschitz, then, by [15, Proposition 6.3], there is MM such that ϕ+M\phi+M is the profile of the unit ball of a homogeneous distance dd in \mathbb{H}. If v2{0}v\in\mathbb{R}^{2}\setminus\{0\}, then there is a sequence {tn}n\{t_{n}\}_{n\in\mathbb{N}} with tn+t_{n}\to+\infty such that f(|v|Mtn)=0f(\frac{|v|\sqrt{M}}{\sqrt{t_{n}}})=0, i.e., de((vMtn,M))=1=de((0,M))d_{e}((\frac{v\sqrt{M}}{\sqrt{t_{n}}},M))=1=d_{e}((0,M)). Therefore,

de((v,tn))de((0,tn))+de((v,0))=tnM(de((vMtn),M))de((0,M)))+de((v,0))=de((v,0)),d_{e}((v,t_{n}))-d_{e}((0,t_{n}))+d_{e}((v,0))=\sqrt{\frac{t_{n}}{M}}\left(d_{e}((\tfrac{v\sqrt{M}}{\sqrt{t_{n}}}),M))-d_{e}((0,M))\right)+d_{e}((v,0))=d_{e}((v,0)),

for all nn\in\mathbb{N}. We conclude that (15) cannot hold for such dd.

5. Horofunctions in polygonal sub-Finsler metrics on \mathbb{H}

Before stating the main result of the section and the paper, we introduce the necessary notation for the description of sub-Finsler distances in \mathbb{H}.

5.1. Geometry of polygonal sub-Finsler metrics

On 2\mathbb{R}^{2}, we denote by ,\langle\cdot,\cdot\rangle the standard scalar product, and by JJ the “multiplication by ii”, i.e., the anticlockwise rotation by π2\frac{\pi}{2}. Notice that ω(,)=J,\omega(\cdot,\cdot)=\langle J\cdot,\cdot\rangle is the standard symplectic form. We will use the symplectic duality between 2\mathbb{R}^{2} and (2)(\mathbb{R}^{2})^{*} induced by ω\omega via

2αωα=ω(αω,)(2).\mathbb{R}^{2}\ni\alpha^{\omega}\leftrightarrow\alpha=\omega(\alpha^{\omega},\cdot)\in(\mathbb{R}^{2})^{*}.

Let QQ be a centrally-symmetric polygon in 2\mathbb{R}^{2} with 2N2N vertices, and let \|\cdot\| be the norm on 2\mathbb{R}^{2} with unit metric disk QQ. Enumerate the vertices {𝗏k}k\{\mathsf{v}_{k}\}_{k} of QQ with kk\in\mathbb{Z} modulo 2N2N, in an anticlockwise order. Notice that 𝗏k=𝗏k+N-\mathsf{v}_{k}=\mathsf{v}_{k+N}. Define the kk-th edge to be the vector ek:=𝗏k+1𝗏ke_{k}:=\mathsf{v}_{k+1}-\mathsf{v}_{k}. For each kk, let αk(2)\alpha_{k}\in(\mathbb{R}^{2})^{*} be the linear function such that αk(𝗏k+tek)=1\alpha_{k}(\mathsf{v}_{k}+te_{k})=1 for all tt\in\mathbb{R}, that is,

αk=ω(ek,)ω(ek,𝗏k),\alpha_{k}=\frac{\omega(e_{k},\cdot)}{\omega(e_{k},\mathsf{v}_{k})},

where ω(ek,𝗏k)=ω(𝗏k+1,𝗏k)0\omega(e_{k},\mathsf{v}_{k})=\omega(\mathsf{v}_{k+1},\mathsf{v}_{k})\neq 0.

Let Q(2)Q^{*}\subset(\mathbb{R}^{2})^{*} be the unit disk of the norm dual to \|\cdot\|, that is, the polar dual of QQ. Note that QQ^{*} is the polygon with vertices {αk}k\{\alpha_{k}\}_{k}.

A result of Busemann [3] tells us that the isoperimetric set II, or isoperimetrix, in (2,)(\mathbb{R}^{2},\|\cdot\|) is the image of QQ^{*} in 2\mathbb{R}^{2} via the symplectic duality.111In other words, II is JQJ^{*}Q^{*}, seen as a subset of 2\mathbb{R}^{2} via the equivalence 2(2)\mathbb{R}^{2}\simeq(\mathbb{R}^{2})^{*} given by the scalar product. It follows that II is the polygon with vertices

αkω=ekω(ek,𝗏k),\alpha_{k}^{\omega}=\frac{e_{k}}{\omega(e_{k},\mathsf{v}_{k})},

where ω(ek,𝗏k)=ω(𝗏k+1,𝗏k)<0\omega(e_{k},\mathsf{v}_{k})=\omega(\mathsf{v}_{k+1},\mathsf{v}_{k})<0, and edges

σk:=αkωαk1ω.\sigma_{k}:=\alpha_{k}^{\omega}-\alpha_{k-1}^{\omega}.

Figure 3 describes the situation for a hexagonal QQ. We note that σk\sigma_{k} is a scalar multiple of 𝗏k\mathsf{v}_{k},

(18) σk=σk𝗏k where σk=αk(σk)>0.\sigma_{k}=\|\sigma_{k}\|\mathsf{v}_{k}\text{ where }\|\sigma_{k}\|=\alpha_{k}(\sigma_{k})>0.

Indeed, ω(σk,𝗏k)=αk(𝗏k)αk1(𝗏k)=11=0\omega(\sigma_{k},\mathsf{v}_{k})=\alpha_{k}(\mathsf{v}_{k})-\alpha_{k-1}(\mathsf{v}_{k})=1-1=0, and thus σk=αk(σk)𝗏k\sigma_{k}=\alpha_{k}(\sigma_{k})\mathsf{v}_{k}. Since αk(σk)=ω(αk1ω,αkω)>0\alpha_{k}(\sigma_{k})=\omega(\alpha_{k-1}^{\omega},\alpha_{k}^{\omega})>0, we have σk=αk(σk)\|\sigma_{k}\|=\alpha_{k}(\sigma_{k}).

𝗏1\mathsf{v}_{1}𝗏2\mathsf{v}_{2}𝗏3\mathsf{v}_{3}𝗏4\mathsf{v}_{4}𝗏5\mathsf{v}_{5}𝗏6\mathsf{v}_{6}e1e_{1}e2e_{2}e3e_{3}e4e_{4}e5e_{5}e6e_{6}QQα1\alpha_{1}α2\alpha_{2}α3\alpha_{3}α4\alpha_{4}α5\alpha_{5}α6\alpha_{6}QQ^{*}σ3\sigma_{3}σ4\sigma_{4}σ5\sigma_{5}σ6\sigma_{6}σ1\sigma_{1}σ2\sigma_{2}IIα2ω\alpha^{\omega}_{2}α4ω\alpha^{\omega}_{4}α3ω\alpha^{\omega}_{3}α5ω\alpha^{\omega}_{5}α1ω\alpha^{\omega}_{1}α6ω\alpha^{\omega}_{6}
Figure 3. Example of a norm ball, dual ball, and isoperimetrix.

For the case of polygonal sub-Finsler metrics on \mathbb{H}, Duchin–Mooney [6] classify geodesics and describe the shape of the unit sphere. Here, we introduce some of their notation and summarize some key results.

Duchin–Mooney break geodesics into two categories: beelines and trace paths. Beeline geodesics are lifts of (2,)(\mathbb{R}^{2},\lVert\cdot\rVert)-geodesics to admissible paths in \mathbb{H}. Trace path geodesics, on the other hand, are lifts of paths in the plane which trace some portion of the boundary of rescaled versions of II.

As in Duchin–Mooney, we partition QQ into quadrilateral regions which are reached by trace paths which trace the same edges of II. That is, for i<j<2N+ii<j<2N+i, define Qij2Q_{ij}\subset\mathbb{R}^{2} to be the set of all endpoints of positively-oriented trace paths in the plane whose parametrizations start by tracing a portion of σi\sigma_{i}, trace all of σi+1,,σj1\sigma_{i+1},\ldots,\sigma_{j-1}, and end by tracing a portion of σj\sigma_{j}, rescaled so that the total length is 1:

Qij\displaystyle Q_{ij} :={1μ(rσi+σi+1++σj1+sσj):r,s[0,1]},\displaystyle:=\left\{\frac{1}{\mu}(r\sigma_{i}+{\sigma_{i+1}}+\ldots+{\sigma_{j-1}}+s{\sigma_{j}}):r,s\in[0,1]\right\},

where μ=μij(r,s)=rσi+σi+1++σj1+sσj\mu=\mu_{ij}(r,s)=r\lVert\sigma_{i}\rVert+\lVert\sigma_{i+1}\rVert+\ldots+\lVert\sigma_{j-1}\rVert+s\lVert\sigma_{j}\rVert normalizes the length of the path. The case i=ji=j is similarly defined

Qii:={1μ(rσi+σi+1++σi1+sσi):r,s[0,1],r+s1}{1μrσi:r[0,1]},Q_{ii}:=\left\{\frac{1}{\mu}(r\sigma_{i}+{\sigma_{i+1}}+\ldots+{\sigma_{i-1}}+s{\sigma_{i}}):r,s\in[0,1],\ r+s\leq 1\right\}\bigcup\left\{\frac{1}{\mu}r\sigma_{i}:r\in[0,1]\right\},

applying the convention that indices are modulo 2N2N. Note that Qii={s𝗏i:s[0,A(I1)]}{𝗏i}Q_{ii}=\{-s\mathsf{v}_{i}:s\in[0,A(I_{1})]\}\cup\{\mathsf{v}_{i}\}, where A(I1)A(I_{1}) is the area of the unit-perimeter scaled copy of II. Then Qi,i+1={s𝗏i+(1s)𝗏i+1:s[0,1]}Q_{i,i+1}=\{s\mathsf{v}_{i}+(1-s)\mathsf{v}_{i+1}:s\in[0,1]\} is the ii-th edge of QQ (see [6, Theorem 7]). For j{i,i+1}j\notin\{i,i+1\},

the regions QijQ_{ij} are non-degenerate quadrilaterals with disjoint interiors, and the set of all QijQ_{ij} covers QQ.

The unit sphere of a polygonal sub-Finsler distance is the set of all endpoints of unit-length geodesics and it can be described as a the region between the graphs of two functions QQ\to\mathbb{R}, see Figure 4. Endpoints of beeline geodesics make up vertical wall panels on the edges of QQ: we denote by Paneli,i+1\mathrm{Panel}_{i,i+1} the vertical wall panel which projects to edge Qi,i+1Q_{i,i+1}, through vertices 𝗏i\mathsf{v}_{i} and 𝗏i+1\mathsf{v}_{i+1}.

Endpoints of all unit-length, positively-oriented trace path geodesics make up the ceiling of the sphere: we denote by Panelij+\mathrm{Panel}_{ij}^{+} the ceiling panel above QijQ_{ij}, that is the set of endpoints of lifts of all unit-length, positively-oriented trace paths whose endpoints lie in QijQ_{ij}.

Refer to captionRefer to captionRefer to caption
Figure 4. On the left and right, two views of a unit sphere with vertical walls missing. In the center, the norm ball QQ broken up into quadrilaterals QijQ_{ij} of points reached by trace paths of the same shape. Figures adapted from [6].

It will be useful to have an explicit description of these panels. Fix a non-degenerate quadrilateral region QijQ_{ij} and define u=uij:[0,1]2Qiju=u_{ij}:[0,1]^{2}\to Q_{ij} by

(19) u(r,s)=(rσi+σij+sσj)μ,u(r,s)=\frac{(r\sigma_{i}+\sigma_{ij}+s{\sigma_{j}})}{\mu},

where σij:=i<k<jσk\sigma_{ij}:=\sum_{i<k<j}\sigma_{k} and μ\mu is again rσi+σi+1++σj1+sσjr\lVert\sigma_{i}\rVert+\lVert\sigma_{i+1}\rVert+\ldots+\lVert\sigma_{j-1}\rVert+s\lVert\sigma_{j}\rVert. Given (r,s)[0,1]2(r,s)\in[0,1]^{2}, the point u(r,s)u(r,s) is the endpoint of a trace path ending in QijQ_{ij}. We then define ϕ=ϕij:[0,1]2[0,+)\phi=\phi_{ij}:[0,1]^{2}\to[0,+\infty) to be the height of the geodesic lift of the trace path implicitly described by u(r,s)u(r,s); that is ϕ(r,s)\phi(r,s) is the balayage area spanned by the curve: 222 To help the reader, we remark that, if γ:[0,1]2\gamma:[0,1]\to\mathbb{R}^{2} is a piecewise affine curve tracing the vectors u0,u1,,un2u_{0},u_{1},\dots,u_{n}\in\mathbb{R}^{2}, then the balayage area spanned by γ\gamma is 120a<bnω(ua,ub).\frac{1}{2}\sum_{0\leq a<b\leq n}\omega(u_{a},u_{b}).

(20) ϕ(r,s)=12μ2(i<k<jω(rσi,σk)+ω(rσi,sσj)+i<k1<k2<jω(σk1,σk2)+i<k<jω(σk,sσj)).\phi(r,s)=\frac{1}{2\mu^{2}}\left(\sum_{i<k<j}\omega(r\sigma_{i},\sigma_{k})+\omega(r\sigma_{i},s\sigma_{j})+\sum_{i<k_{1}<k_{2}<j}\omega(\sigma_{k_{1}},\sigma_{k_{2}})+\sum_{i<k<j}\omega(\sigma_{k},s\sigma_{j})\right).

The map [0,1]2[0,1]^{2}\to\mathbb{H}, (r,s)(uij(r,s),ϕij(r,s))(r,s)\mapsto(u_{ij}(r,s),\phi_{ij}(r,s)) is then a parametrization of Panelij+\mathrm{Panel}^{+}_{ij}. Similarly, the map [0,1]2[0,1]^{2}\to\mathbb{H}, (r,s)(uij(r,s),ϕij(r,s))(r,s)\mapsto(u_{ij}(r,s),-\phi_{ij}(r,s)) is a parametrization of Panelij\mathrm{Panel}^{-}_{ij}, the part of the basement of B\partial B that projects onto QijQ_{ij}, see [6, Theorem 10]. Note that this parametrization of the basement panel does not encode the combinatorial information of the trace path geodesic that leads there.

Remark 5.1.

We will prove most of our results for ceiling points. The basement case is then derived via the involutive automorphism Θ:\Theta:\mathbb{H}\to\mathbb{H}, Θ(v,t)=(Θ(v),t)\Theta(v,t)=(\Theta(v),-t), where Θ:22\Theta:\mathbb{R}^{2}\to\mathbb{R}^{2} is the linear transformation that maps 𝗏2N\mathsf{v}_{2N} to itself and flips the line orthogonal to 𝗏2N\mathsf{v}_{2N}. Notice that Θ2=Id\Theta^{2}=\mathrm{Id} and Θω=ω\Theta^{*}\omega=-\omega.

The basement of the unit ball BB of dd is mapped to the ceiling of the unit ball BΘB^{\Theta} of the new distance dΘ(x,y):=d(Θ(x),Θ(y))d^{\Theta}(x,y):=d(\Theta(x),\Theta(y)). The distance dΘd^{\Theta} is again sub-Finsler with unit disk QΘ=Θ(Q)Q^{\Theta}=\Theta(Q). The polygon QΘQ^{\Theta} has vertices 𝗏kΘ:=Θ(𝗏k)\mathsf{v}^{\Theta}_{k}:=\Theta(\mathsf{v}_{-k}), in anti-clockwise order, and edges ekΘ=Θ(ek1)e^{\Theta}_{k}=\Theta(-e_{-k-1}). It follows that

αkΘΘ=αk1,\alpha^{\Theta}_{k}\circ\Theta=-\alpha_{-k-1},

hence αkΘ,ω=Θ(αk1ω)\alpha^{\Theta,\omega}_{k}=\Theta(\alpha^{\omega}_{-k-1}) and σkΘ=Θ(σk)\sigma^{\Theta}_{k}=-\Theta(\sigma_{-k}) with σkΘΘ=σk\|\sigma^{\Theta}_{k}\|^{\Theta}=\|\sigma_{-k}\|. See Figure 5. Therefore, if i<ji<j, then

uj,iΘ(s,r)=Θ(uij(r,s)).u_{-j,-i}^{\Theta}(s,r)=-\Theta(u_{ij}(r,s)).

So, if pBp\in\partial B lies in the basement and π(p)=uij(r,s)\pi(p)=u_{ij}(r,s), then Θ(p)BΘ\Theta(p)\in\partial B^{\Theta} lies in the ceiling and π(Θ(p))=uj,iΘ(s,r)=uj+N,i+NΘ(s,r)\pi(\Theta(p))=-u_{-j,-i}^{\Theta}(s,r)=u^{\Theta}_{-j+N,-i+N}(s,r). See Figure 6. Finally, for all p,vp,v\in\mathbb{H}, we have

P

​​D
deΘ
|
Θ(p)
[Θ(v)]
=

P

​​D
de
|
p
[v]
.
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d^{\Theta}_{e}|_{\Theta(p)}[\Theta(v)]=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}[v].
𝗏1\mathsf{v}_{1}𝗏2\mathsf{v}_{2}𝗏3\mathsf{v}_{3}𝗏4\mathsf{v}_{4}𝗏5\mathsf{v}_{5}𝗏6\mathsf{v}_{6}e1e_{1}e2e_{2}e3e_{3}e4e_{4}e5e_{5}e6e_{6}QQσ3\sigma_{3}σ4\sigma_{4}σ5\sigma_{5}σ6\sigma_{6}σ1\sigma_{1}σ2\sigma_{2}IIα2ω\alpha^{\omega}_{2}α4ω\alpha^{\omega}_{4}α3ω\alpha^{\omega}_{3}α5ω\alpha^{\omega}_{5}α1ω\alpha^{\omega}_{1}α6ω\alpha^{\omega}_{6}Θ\ThetaΘ\Theta𝗏5Θ\mathsf{v}^{\Theta}_{5}𝗏4Θ\mathsf{v}^{\Theta}_{4}𝗏3Θ\mathsf{v}^{\Theta}_{3}𝗏2Θ\mathsf{v}^{\Theta}_{2}𝗏1Θ\mathsf{v}^{\Theta}_{1}𝗏6Θ\mathsf{v}^{\Theta}_{6}e5Θe^{\Theta}_{5}e4Θe^{\Theta}_{4}e3Θe^{\Theta}_{3}e2Θe^{\Theta}_{2}e1Θe^{\Theta}_{1}e6Θe^{\Theta}_{6}Θ(Q)\Theta(Q)σ3Θ\sigma^{\Theta}_{3}σ2Θ\sigma^{\Theta}_{2}σ1Θ\sigma^{\Theta}_{1}σ6Θ\sigma^{\Theta}_{6}σ5Θ\sigma^{\Theta}_{5}σ4Θ\sigma^{\Theta}_{4}IΘI^{\Theta}α3Θ,ω\alpha^{\Theta,\omega}_{3}α2Θ,ω\alpha^{\Theta,\omega}_{2}α1Θ,ω\alpha^{\Theta,\omega}_{1}α6Θ,ω\alpha^{\Theta,\omega}_{6}α5Θ,ω\alpha^{\Theta,\omega}_{5}α4Θ,ω\alpha^{\Theta,\omega}_{4}
Figure 5. Example of a norm ball and isoperimetrix, with their transformations under Θ\Theta.
rσ1r\sigma_{1}σ2\sigma_{2}σ3\sigma_{3}σ4\sigma_{4}sσ5s\sigma_{5}sσ5-s\sigma_{5}σ4-\sigma_{4}σ3-\sigma_{3}σ2-\sigma_{2}rσ1-r\sigma_{1}rΘ(σ1)=rσ1Θr\Theta(\sigma_{1})=r\sigma^{\Theta}_{-1}σ2Θ\sigma^{\Theta}_{-2}σ3Θ\sigma^{\Theta}_{-3}σ4Θ\sigma^{\Theta}_{-4}sσ5Θs\sigma^{\Theta}_{-5}
Figure 6. Three trace paths with similar combinatorics whose lifts end at ceiling point p=(u,ϕ(u))p=(u,\phi(u)), basement point p1=(u,ϕ(u))p^{-1}=(-u,-\phi(u)), and the image Θ(p)\Theta(p) of pp under the involution Θ\Theta, respectively. Note that the trace path of p1p^{-1} has the reverse parametrization as that of pp.

5.2. The theorem

Let dd be a polygonal sub-Finsler metric on the Heisenberg group \mathbb{H}. The fundamental lemma identifying horofunctions with Pansu derivatives (Lemma 2.3) applies in this case, but we need to take care in describing all possible blow-ups of the distance function at points on the sphere.

These blow-ups take two forms. As we will explain below, \mathbb{H} is partitioned so that the function ded_{e} is CC^{\infty} in the interior of each region. So, on the one hand, we have the points of the unit sphere where ded_{e} is smooth, and thus the only blow-up is the Pansu derivative of ded_{e}. On the other hand, on the non-smooth part of the unit sphere, which we call the seam, the blow-up of ded_{e} is defined piecewise as in Theorem 3.9.

The unit sphere of dd is made of smooth and non-smooth points. Smooth points are the interior points of the panels on ceiling, basement, and walls. Non-smooth points are on the seams between those panels, that is: north and south poles, star-like seams near the north and south poles, and seams between ceiling or basement and wall panels, vertices of QQ. See Figure 7 for the seams along a hexagonal unit sphere: each type of seam point intersects different combinations of panel dilation cones and hence provides a different kind of blow-up function. We will study the blow-ups of the distance function ded_{e} in each case separately. The results are summarized in the following theorem.

Refer to caption11334422
Figure 7. A top-down view of the four types of seams, in red, on a hexagonal unit sphere; 1) north and south pole; 2) vertices; 3) star-like seams near poles; 4) wall seams.
Theorem 5.2 (Blow-ups of ded_{e}).

Using the notation of Section 5.1, the blow-ups of ded_{e} at a point pp on the sphere of dd fall in one of the following cases.

Smooth points:

  1. (SS1)

    If pp is in the interior of Panelij+\mathrm{Panel}^{+}_{ij} such that the π(p)=uij(r,s)\pi(p)=u_{ij}(r,s), then the Pansu derivative of ded_{e} exists at pp and

    P

    ​​D
    de
    |
    p
    (v,t)
    =((1s)αj1+sαj)(v)
    .
    \text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}(v,t)=((1-s)\alpha_{j-1}+s\alpha_{j})(v).
  2. (SS2)

    If pp is in the interior of Panelij\mathrm{Panel}^{-}_{ij} such that the π(p)=uij(r,s)\pi(p)=u_{ij}(r,s), then the Pansu derivative of ded_{e} exists at pp and

    P

    ​​D
    de
    |
    p
    (v,t)
    =((1r)αi+rαi1)(v)
    .
    \text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}(v,t)=((1-r)\alpha_{i}+r\alpha_{i-1})(v).
  3. (SS3)

    If pp is in the interior of Paneli,i+1\mathrm{Panel}_{i,i+1}, then

    P

    ​​D
    de
    |
    p
    (v,t)
    =αi(v)
    .
    \text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}(v,t)=\alpha_{i}(v).

Non-smooth points:

  1. (S\not{}\!S1)

    North and south poles

    1. (a)

      For w2w\in\mathbb{R}^{2},

      f(v,t)=wwv;f(v,t)=\lVert w\rVert-\lVert w-v\rVert;
    2. (b)

      For CC\in\mathbb{R} and i{1,,2N}i\in\{1,\ldots,2N\}

      f(v,t)={αi(v)+c1ω(𝗏i,v)Cαi1(v)+c2ω(𝗏i,v)>C;f(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\leq C\\ \alpha_{i-1}(v)+c_{2}&\omega(\mathsf{v}_{i},v)>C\end{cases};
    3. (c)

      For i{1,,2N}i\in\{1,\ldots,2N\} (corresponding to C{,+}C\in\{-\infty,+\infty\})

      f(v,t)=αi(v);f(v,t)=\alpha_{i}(v);
  2. (S\not{}\!S2)

    ii-th vertex of QQ, i{1,,2N}i\in\{1,\ldots,2N\}, for C{,+}C\in\mathbb{R}\cup\{-\infty,+\infty\}:

    f(v,t)={αi(v)+c1ω(𝗏i,v)Cαi1(v)+c2ω(𝗏i,v)<C;f(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\geq C\\ \alpha_{i-1}(v)+c_{2}&\omega(\mathsf{v}_{i},v)<C\end{cases};
  3. (S\not{}\!S3)

    Star-like seams

    1. (a)

      Near the north pole, for C{,+}C\in\mathbb{R}\cup\{-\infty,+\infty\} and s(0,1]s\in(0,1]:

      f(v,t)={αi1(v)+c1ω(𝗏i,v)C((1s)αi1+sαi)(v)+c2ω(𝗏i,v)<Cf(v,t)=\begin{cases}\alpha_{i-1}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\geq C\\ ((1-s)\alpha_{i-1}+s\alpha_{i})(v)+c_{2}&\omega(\mathsf{v}_{i},v)<C\end{cases}
    2. (b)

      Near the south pole, for C{,+}C\in\mathbb{R}\cup\{-\infty,+\infty\} and s(0,1]s\in(0,1]:

      f(v,t)={αi(v)+c1ω(𝗏i,v)C((1s)αi+sαi1)(v)+c2ω(𝗏i,v)>Cf(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\leq C\\ ((1-s)\alpha_{i}+s\alpha_{i-1})(v)+c_{2}&\omega(\mathsf{v}_{i},v)>C\end{cases}
  4. (S\not{}\!S4)

    Wall seams

    1. (a)

      Between wall and ceiling panels, for C{,+}C\in\mathbb{R}\cup\{-\infty,+\infty\} and s(0,1]s\in(0,1]:

      f(v,t)={αi1(v)+c1ω(𝗏i,v)C((1s)αi1+sαi)(v)+c2ω(𝗏i,v)>Cf(v,t)=\begin{cases}\alpha_{i-1}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\leq C\\ ((1-s)\alpha_{i-1}+s\alpha_{i})(v)+c_{2}&\omega(\mathsf{v}_{i},v)>C\end{cases}
    2. (b)

      Between wall and basement panels, for C{,+}C\in\mathbb{R}\cup\{-\infty,+\infty\} and s(0,1]s\in(0,1]:

      f(v,t)={αi(v)+c1ω(𝗏i,v)C((1s)αi+sαi1)(v)+c2ω(𝗏i,v)<Cf(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\geq C\\ ((1-s)\alpha_{i}+s\alpha_{i-1})(v)+c_{2}&\omega(\mathsf{v}_{i},v)<C\end{cases}

We remark that in each of the non-smooth cases, the constants c1c_{1} and c2c_{2} are uniquely determined by the value of CC since by definition f(0,0)=0f(0,0)=0 and ff is continuous. The proof of this theorem is the content of the rest of the section. Before diving into it, we present several consequences, in particular the description of the horoboundary of (,d)(\mathbb{H},d).

Theorem 5.2 has corollaries concerning the regularity of ded_{e} on the sphere. Indeed, since the Pansu derivative of ded_{e} on the ceiling depends only on the endpoints of trace paths, it follows that

P

​​D
de
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}
is continuous on the ceiling and the basement of B\partial B, except to the star-like seams near the poles, in red in Figure 8. We could draw a similar figure for the basement, where the families would spiral in anticlockwise, instead of clockwise.

Corollary 5.3.

Except for star-like sets near the north and south poles, the function ded_{e} has continuous Pansu derivative

in the interior of the ceiling and the basement of B\partial B.

Corollary 5.3 with Proposition 2.1 implies that B\partial B is a CH1C^{1}_{H} submanifold of \mathbb{H} (in the sense of [10]) at every interior point of the ceiling or basement, outside the star-like sets.

Refer to caption
Refer to caption
Figure 8. A top-down view of families of ceiling points (left) and a bottom-up view of families of basement points (right) with the same Pansu derivatives.
Theorem 5.4 (The horofunction boundary of (,d)(\mathbb{H},d)).

The horofunction boundary of (,d)(\mathbb{H},d) is the union of the image of the following embeddings in C()C(\mathbb{H}): First, a disk given by K:2C()K:\mathbb{R}^{2}\to C(\mathbb{H}), wfww\mapsto f_{w}, where fw(v,s)=wvwf_{w}(v,s)=\|w\|-\|v-w\|. The boundary of K(2)K(\mathbb{R}^{2}) in C()C(\mathbb{H}) is h(2,)-\partial_{h}(\mathbb{R}^{2},\lVert\cdot\rVert).

Second, for each ii, we have the following maps [0,1]×[,]C()[0,1]\times[-\infty,\infty]\to C(\mathbb{H}):

ψi(s,a)\displaystyle\psi_{i}^{\vee}(s,a) =(αi1(a0))((1s)αi1+sαi+(a0))\displaystyle=\left(\alpha_{i-1}-(a\vee 0)\right)\vee\left((1-s)\alpha_{i-1}+s\alpha_{i}+(a\wedge 0)\right)
ψi(s,a)\displaystyle\psi_{i}^{\wedge}(s,a) =(αi1(a0))((1s)αi1+sαi+(a0))\displaystyle=\left(\alpha_{i-1}-(a\wedge 0)\right)\wedge\left((1-s)\alpha_{i-1}+s\alpha_{i}+(a\vee 0)\right)
ξi(s,a)\displaystyle\xi_{i}^{\vee}(s,a) =(αi(a0))((1s)αi1+sαi+(a0))\displaystyle=\left(\alpha_{i}-(a\vee 0)\right)\vee\left((1-s)\alpha_{i-1}+s\alpha_{i}+(a\wedge 0)\right)
ξi(s,a)\displaystyle\xi_{i}^{\wedge}(s,a) =(αi(a0))((1s)αi1+sαi+(a0)).\displaystyle=\left(\alpha_{i}-(a\wedge 0)\right)\wedge\left((1-s)\alpha_{i-1}+s\alpha_{i}+(a\vee 0)\right).

For each ii, the image of these four maps is two spheres glued together along a meridian. The second meridian of each of the two spheres is the segment between αi1\alpha_{i-1} and αi\alpha_{i} in h(2,)\partial_{h}(\mathbb{R}^{2},\|\cdot\|) and h(2,)-\partial_{h}(\mathbb{R}^{2},\|\cdot\|), respectively.

Theorem 5.4 is proven by inspection of the functions listed in Theorem 5.2.

It turns out that all of the smooth points on the ceiling, basement, and vertical walls of B\partial B contribute a circle’s worth of functions to the horoboundary. Indeed, they all have Pansu derivatives which lie in LL^{*}, the boundary of the dual ball QQ^{*}. This is analogous to results in the sub-Riemannian case; Klein–Nicas in [13] showed that the smooth points contribute a circle’s worth of functions to the boundary, while the rest of the boundary comes from vertical sequences, analogous to our Theorem 4.1.

See Figures 10 and 9. In Figure 10, we introduce a sense of directionality to the horofunction boundary. Recall that to any sequence {qn}\{q_{n}\}\subset\mathbb{H} converging to a horofunction, we can associate sequences {pn}nB\{p_{n}\}_{n}\subset\partial B and {ϵn}n\{\epsilon_{n}\}_{n}\subset\mathbb{R}, where δϵnqn=pn\delta_{\epsilon_{n}}q_{n}=p_{n}. For each horofunction fhf\in\partial_{h}\mathbb{H}, there exist sequences {qn}n({pn}n,{ϵn}n)\{q_{n}\}_{n}\leftrightarrow(\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) such that qnfq_{n}\to f and pnpBp_{n}\to p\in\partial B. This assigns directions to horofunctions in the boundary. This correspondence between the boundary and the unit sphere is far from bijective. There exist families of directions, such as each blue vertical wall panel, which collapse to single points in the boundary. On the other hand, there are directions, such as the purple north and south poles, which blow-up to 1- or 2- dimensional families in the boundary. In these cases, which boundary point you converge to will depend on how exactly qnq_{n} goes off to infinity. The colors in the figures allow us to see which directions on the sphere converge to which families horofunctions.

Corollary 5.5.

Let dd be a polygonal sub-Finsler metric on \mathbb{H}. Then the set of Busemann functions is homeomorphic to a circle.

Proof.

The only infinite geodesic rays based at the origin are the beeline geodesics, the lifts of LL-norm geodesics in the plane to admissible paths in \mathbb{H}, as described in Section 5.1. Thus, the set of Busemann functions comes from blow-ups of points in the vertical walls and vertices of the unit sphere and is isomorphic to h(2,Q)S1\partial_{h}(\mathbb{R}^{2},\|\cdot\|_{Q})\leavevmode\nobreak\ \cong\leavevmode\nobreak\ S^{1}. ∎

ψi\psi^{\wedge}_{i}ψi\psi^{\vee}_{i}0-\inftyss11\infty(αi1(a0))(αi+(a0))(\alpha_{i-1}-(a\wedge 0))\wedge(\alpha_{i}+(a\vee 0))αi1\alpha_{i-1}aaαi\alpha_{i}0-\inftyss11\infty(αi1(a0))(αi+(a0))(\alpha_{i-1}-(a\vee 0))\vee(\alpha_{i}+(a\wedge 0))(1s)αi1+sαi(1-s)\alpha_{i-1}+s\alpha_{i}aaαi\alpha_{i}αi\alpha_{i}αi1\alpha_{i-1}ξi\xi^{\wedge}_{i}ξi\xi^{\vee}_{i}0-\inftyss11\inftyaa0-\inftyss11\inftyaaαi\alpha_{i}αi1\alpha_{i-1}
Figure 9. A schematic description of the maps ψi\psi_{i}^{\vee}, ψi\psi_{i}^{\wedge}, ξi\xi_{i}^{\vee} and ξi\xi_{i}^{\wedge}.
Refer to caption
Refer to caption
Figure 10. The unit sphere coming from the hexagonal norm on the left, and the corresponding sub-Finsler boundary on the right.

5.3. Blow-ups of ded_{e} at smooth points

First we consider blow-ups of ded_{e} at smooth points on B\partial B, such as in the interior of each ceiling, basement, or wall panel making up the unit sphere. Since ded_{e} is smooth in the interior of each of these panels, it is strictly Pansu differentiable.

We know from above that ceiling and basement points are reached by geodesics which are lifts of trace paths. It turns out that the Pansu derivative of ded_{e} on the ceiling or basement depends only on where in the isoperimetrix II the trace path ends and is independent of the rest of the shape of the trace path.

Proposition 5.6 (Ceiling and basement Pansu derivatives).

If pBp\in\partial B is a ceiling point with π(p)=uij(r,s)\pi(p)=u_{ij}(r,s), j{i,i+1}j\notin\{i,i+1\},

then the Pansu derivative of ded_{e} exists at pp, and

P

​​D
de
|
p
(v,t)
=((1s)αj1+sαj)(v)
.
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}(v,t)=((1-s)\alpha_{j-1}+s\alpha_{j})(v).

Similarly, if pBp\in\partial B is a basement point with π(p)=uij(r,s)\pi(p)=-u_{ij}(r,s), i<ji<j, then the Pansu derivative of ded_{e} exists at pp, and

P

​​D
de
|
p
(v,t)
=((1r)αi+rαi1)(v)
.
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}(v,t)=((1-r)\alpha_{i}+r\alpha_{i-1})(v).
Proof.

Given that pp is in the interior of Panelij+\mathrm{Panel}^{+}_{ij}, ded_{e} is smooth at pp, and hence the Pansu derivative exists. Pansu derivatives are linear and invariant on vertical fibers, so we are looking for a linear functional A(2)A\in(\mathbb{R}^{2})^{*} such that

P

​​D
de(p)(v,t)
=A[v]
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}(p)(v,t)=A[v]
.

Let γ1:[0,1+ϵ]\gamma_{1}:[0,1+\epsilon]\to\mathbb{H} be the unit-speed trace path geodesic from the origin to γ1(1)=p\gamma_{1}(1)=p. Since pp is in the interior of Panelij+\mathrm{Panel}_{ij}^{+},

for sufficiently small hh we have γ1(1+h)=pδh𝗏j\gamma_{1}(1+h)=p\delta_{h}\mathsf{v}_{j}, and so

1=limh0de(γ1(1+h))de(γ1(1))h=limh0de(pδh𝗏j)de(p)h=

P

​​D
de
|
p
(𝗏j)
.
1=\lim_{h\to 0}\frac{d_{e}(\gamma_{1}(1+h))-d_{e}(\gamma_{1}(1))}{h}=\lim_{h\to 0}\frac{d_{e}(p\delta_{h}\mathsf{v}_{j})-d_{e}(p)}{h}=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}(\mathsf{v}_{j}).

Next, let γ2:(ϵ,ϵ)\gamma_{2}:(-\epsilon,\epsilon)\to\mathbb{H} be a C2C^{2} path along the unit sphere B\partial B which is horizontal at pp, with γ2(0)=p\gamma_{2}(0)=p and γ2(0)=wΔp\gamma_{2}^{\prime}(0)=w\in\Delta_{p}. Since γ2\gamma_{2} is on the unit sphere, de(γ2)1d_{e}(\gamma_{2})\equiv 1. A consequence of the horizontality of γ2\gamma_{2} at pp is the limit limh0δ1/h(p1γ2(h))=γ2(0)=w\lim_{h\to 0}\delta_{1/h}(p^{-1}\gamma_{2}(h))=\gamma_{2}^{\prime}(0)=w, and so by the Pansu differentiability of ded_{e} at pp,

0=limh0de(γ2(h))de(γ2(0))h=

P

​​D
de
|
p
(w)
.
0=\lim_{h\to 0}\frac{d_{e}(\gamma_{2}(h))-d_{e}(\gamma_{2}(0))}{h}=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}(w).

Now, we seek an expression for ww.

Since Panelij+={(u(r,s),ϕ(r,s)):r,s[0,1]}\mathrm{Panel}_{ij}^{+}=\{(u(r,s),\phi(r,s)):r,s\in[0,1]\}, where uu and ϕ\phi are defined in (19) and in (20), the tangent bundle to the unit sphere has a frame given by (rurϕ)\left(\begin{smallmatrix}\partial_{r}u\\ \partial_{r}\phi\end{smallmatrix}\right) and (susϕ)\left(\begin{smallmatrix}\partial_{s}u\\ \partial_{s}\phi\end{smallmatrix}\right). We take the partial derivatives of uu and ϕ\phi

(21) ru(r,s)\displaystyle\partial_{r}u(r,s) =σiμ(𝗏iu(r,s)),\displaystyle=\frac{\lVert\sigma_{i}\rVert}{\mu}(\mathsf{v}_{i}-u(r,s)), rϕ(r,s)=σiμ(12ω(𝗏i,u(r,s)rσiμ)2ϕ(r,s)),\displaystyle\partial_{r}\phi(r,s)=\frac{\lVert\sigma_{i}\rVert}{\mu}\left(\frac{1}{2}\omega(\mathsf{v}_{i},u(r,s)-r\frac{\sigma_{i}}{\mu})-2\phi(r,s)\right),
su(r,s)\displaystyle\partial_{s}u(r,s) =σjμ(𝗏ju(r,s)),\displaystyle=\frac{\lVert\sigma_{j}\rVert}{\mu}(\mathsf{v}_{j}-u(r,s)), sϕ(r,s)=σjμ(12ω(u(r,s)sσjμ,𝗏j)2ϕ(r,s)),\displaystyle\partial_{s}\phi(r,s)=\frac{\lVert\sigma_{j}\rVert}{\mu}\left(\frac{1}{2}\omega(u(r,s)-s\frac{\sigma_{j}}{\mu},\mathsf{v}_{j})-2\phi(r,s)\right),

where again μ=rσi+σi+1++σj1+sσj\mu=r\lVert\sigma_{i}\rVert+\lVert\sigma_{i+1}\rVert+\ldots+\lVert\sigma_{j-1}\rVert+s\lVert\sigma_{j}\rVert. If p=(u,ϕ)p=(u,\phi) has trace coordinates (r,s)(r,s) in Panelij+\mathrm{Panel}_{ij}^{+}, then after rescaling and simplifying, we have

TpB= Span{(𝗏iu(r,s)12ω(𝗏i,u(r,s))2ϕ(r,s)),(𝗏ju(r,s)12ω(u(r,s),𝗏j)2ϕ(r,s))}.\ T_{p}\partial B=\text{ Span}\left\{\begin{pmatrix}\mathsf{v}_{i}-u(r,s)\\ \frac{1}{2}\omega(\mathsf{v}_{i},u(r,s))-2\phi(r,s)\end{pmatrix},\begin{pmatrix}\mathsf{v}_{j}-u(r,s)\\ \frac{1}{2}\omega(u(r,s),\mathsf{v}_{j})-2\phi(r,s)\end{pmatrix}\right\}.

Meanwhile, the horizontal subspace at pp is spanned by the left translations from the origin to pp of (𝗏iu(r,s),0)(\mathsf{v}_{i}-u(r,s),0) and (𝗏ju(r,s),0)(\mathsf{v}_{j}-u(r,s),0).

This gives

Δp= Span{(𝗏iu(r,s)12ω(u(r,s),𝗏i)),(𝗏ju(r,s)12ω(u(r,s),𝗏j))}.\Delta_{p}=\text{ Span}\left\{\begin{pmatrix}\mathsf{v}_{i}-u(r,s)\\ \frac{1}{2}\omega(u(r,s),\mathsf{v}_{i})\end{pmatrix},\begin{pmatrix}\mathsf{v}_{j}-u(r,s)\\ \frac{1}{2}\omega(u(r,s),\mathsf{v}_{j})\end{pmatrix}\right\}.

These two bases for TpBT_{p}\partial B and Δp\Delta_{p} allow us to find ww in the intersection as

(22) w=(2ϕ(r,s)(𝗏j𝗏i)+ω(u,𝗏i)(𝗏ju)ϕ(r,s)ω(u,𝗏j𝗏i)+12ω(u,𝗏i)ω(u,𝗏j)).w=\begin{pmatrix}2\phi(r,s)(\mathsf{v}_{j}-\mathsf{v}_{i})+\omega(u,\mathsf{v}_{i})(\mathsf{v}_{j}-u)\\ \phi(r,s)\omega(u,\mathsf{v}_{j}-\mathsf{v}_{i})+\frac{1}{2}\omega(u,\mathsf{v}_{i})\omega(u,\mathsf{v}_{j})\end{pmatrix}.

The vector ww is the left-translation from the origin to pp of the horizontal vector (w^,0)(\hat{w},0), where w^:=2ϕ(r,s)(𝗏j𝗏i)+ω(u,𝗏i)(𝗏ju)\hat{w}:=2\phi(r,s)(\mathsf{v}_{j}-\mathsf{v}_{i})+\omega(u,\mathsf{v}_{i})(\mathsf{v}_{j}-u). Notice that, if we set A=(1s)αj1+sαjA=(1-s)\alpha_{j-1}+s\alpha_{j}, then in order to prove

P

​​D
de
|
p
=A
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}=A
we only need to show that

(23) A[w^]=0,A[\hat{w}]=0,

because Avj=1Av_{j}=1. The proof of (23) is a long computation, of which we describe the main steps. The strategy is to write A[w^]A[\hat{w}] in terms of the symplectic duals αiω\alpha^{\omega}_{i} of the covectors αi\alpha_{i}. One can easily show that

σj=ω(αj1α,αjω), and 𝗏j=αjαj1ω(αj1,αj).\|\sigma_{j}\|=\omega(\alpha^{\alpha}_{j-1},\alpha^{\omega}_{j}),\quad\text{ and }\quad\mathsf{v}_{j}=\frac{\alpha_{j}-\alpha_{j-1}}{\omega(\alpha_{j-1},\alpha_{j})}.

First of all, we have

(24) A[w^]=1μij2((ω(rσi,σij)+ω(σij,sσj)+ω(rσi,sσj)+i<a<b<jω(σa,σb))A[𝗏j𝗏i]+ω(rσi+σij+sσj,𝗏i)A[μ𝗏jrσiσijsσj]).A[\hat{w}]=\frac{1}{\mu_{ij}^{2}}\Bigg{(}\Bigg{(}\omega(r\sigma_{i},\sigma_{ij})+\omega(\sigma_{ij,s\sigma_{j}})+\omega(r\sigma_{i},s\sigma_{j})+\sum_{i<a<b<j}\omega(\sigma_{a},\sigma_{b})\Bigg{)}A[\mathsf{v}_{j}-\mathsf{v}_{i}]\\ +\omega(r\sigma_{i}+\sigma_{ij}+s\sigma_{j},\mathsf{v}_{i})A[\mu\mathsf{v}_{j}-r\sigma_{i}-\sigma_{ij}-s\sigma_{j}]\Bigg{)}.

Secondly, one can check the following equalities:

A[𝗏j𝗏i]\displaystyle A[\mathsf{v}_{j}-\mathsf{v}_{i}] =1A[αiωαi1ω]ω(αi1,αi),\displaystyle=1-\frac{A[\alpha^{\omega}_{i}-\alpha^{\omega}_{i-1}]}{\omega(\alpha_{i-1},\alpha_{i})},
A[μ𝗏jrσiσijsσj]\displaystyle A[\mu\mathsf{v}_{j}-r\sigma_{i}-\sigma_{ij}-s\sigma_{j}] =A[r(αi1ωαiω)+αiω]+μ,\displaystyle=A[r(\alpha^{\omega}_{i-1}-\alpha^{\omega}_{i})+\alpha^{\omega}_{i}]+\mu,
ω(rσi+σij+sσj,𝗏i)\displaystyle\omega(r\sigma_{i}+\sigma_{ij}+s\sigma_{j},\mathsf{v}_{i}) =A[αiωαi1ω]ω(αi1ω,αiω)1,\displaystyle=\frac{A[\alpha^{\omega}_{i}-\alpha^{\omega}_{i-1}]}{\omega(\alpha^{\omega}_{i-1},\alpha^{\omega}_{i})}-1,
μij\displaystyle\mu_{ij} =rω(αi1ω,αiω)+i<k<jω(αk1ω,αkω)+sω(αj1ω,αjω),\displaystyle=r\omega(\alpha^{\omega}_{i-1},\alpha^{\omega}_{i})+\sum_{i<k<j}\omega(\alpha^{\omega}_{k-1},\alpha^{\omega}_{k})+s\omega(\alpha^{\omega}_{j-1},\alpha^{\omega}_{j}),
i<a<b<jω(σa,σb)\displaystyle\sum_{i<a<b<j}\omega(\sigma_{a},\sigma_{b}) =i<k<j1ω(αk1ω,αkω)+ω(αj2ω,αj1ω)ω(αiω,αj1ω),\displaystyle=\sum_{i<k<j-1}\omega(\alpha^{\omega}_{k-1},\alpha^{\omega}_{k})+\omega(\alpha^{\omega}_{j-2},\alpha^{\omega}_{j-1})-\omega(\alpha^{\omega}_{i},\alpha^{\omega}_{j-1}),
ω(rσi,σij)+ω(σij,sσj)+ω(rσi,sσj)=rω(αiωαi1ω,αj1ωαiω)+s(ω(αj1ω,αjω)ω(αiω,αjω)+ω(αiω,αjω))+rs(ω(αiω,αjω)ω(αiω,αj1ω)ω(αi1ω,αjω)+ω(αi1ω,αj1ω)).\omega(r\sigma_{i},\sigma_{ij})+\omega(\sigma_{ij},s\sigma_{j})+\omega(r\sigma_{i},s\sigma_{j})=r\omega(\alpha^{\omega}_{i}-\alpha^{\omega}_{i-1},\alpha^{\omega}_{j-1}-\alpha^{\omega}_{i})+s(\omega(\alpha^{\omega}_{j-1},\alpha^{\omega}_{j})\\ -\omega(\alpha^{\omega}_{i},\alpha^{\omega}_{j})+\omega(\alpha^{\omega}_{i},\alpha^{\omega}_{j}))+rs(\omega(\alpha^{\omega}_{i},\alpha^{\omega}_{j})-\omega(\alpha^{\omega}_{i},\alpha^{\omega}_{j-1})-\omega(\alpha^{\omega}_{i-1},\alpha^{\omega}_{j})+\omega(\alpha^{\omega}_{i-1},\alpha^{\omega}_{j-1})).

Finally, using these formulas to rewrite (24), one easily finds that A[w^]A[\hat{w}] factorizes into 1μij2(1A[αiωαi1ω]ω(αi1,αi))\frac{1}{\mu_{ij}^{2}}\left(1-\frac{A[\alpha^{\omega}_{i}-\alpha^{\omega}_{i-1}]}{\omega(\alpha_{i-1},\alpha_{i})}\right) and a polynomial of order two in rr and ss. Each coefficient of this polynomial is easily shown to be zero, completing the proof of 23.

To show the result for basement points, we will use Remark 5.1 and the result just proved for ceiling points. Let pBp\in\partial B be in the basement with π(p)=uij(r,s)\pi(p)=u_{ij}(r,s), i<ji<j, so that Θ(p)\Theta(p) lies in the ceiling of BΘ\partial B^{\Theta} and π(Θ(p))=uj,iΘ(s,r)=uj+N,i+NΘ(s,r)\pi(\Theta(p))=-u_{-j,-i}^{\Theta}(s,r)=u_{-j+N,-i+N}^{\Theta}(s,r). Then, for all vv\in\mathbb{H},

P

​​D
de
|
p
[v]
\displaystyle\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}[v]
=

P

​​D
deΘ
|
Θ(p)
[Θ(v)]
\displaystyle=\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d^{\Theta}_{e}|_{\Theta(p)}[\Theta(v)]
=((1r)αi1+NΘ+rαi+NΘ)[Θ(v)]\displaystyle=((1-r)\alpha^{\Theta}_{-i-1+N}+r\alpha^{\Theta}_{-i+N})[\Theta(v)]
=((1r)αi1Θ+rαiΘ)[Θ(v)]\displaystyle=-((1-r)\alpha^{\Theta}_{-i-1}+r\alpha^{\Theta}_{-i})[\Theta(v)]
=((1r)αi+rαi1)(v).\displaystyle=((1-r)\alpha_{i}+r\alpha_{i-1})(v).

This completes the proof.

Proposition 5.7 (Wall Pansu derivatives).

If pp is in the interior of the wall panel Paneli,i+1\mathrm{Panel}_{i,i+1}, then

P

​​D
de
|
p
(v,t)
=αi(v)
.
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}(v,t)=\alpha_{i}(v).
Proof.

Let p=(u,t)p=(u,t^{\prime}) be in the interior of Paneli,i+1\mathrm{Panel}_{i,i+1}, and let q=(v,t)q=(v,t)\in\mathbb{H}. For sufficiently small ϵ>0\epsilon>0, the point pδϵqp\delta_{\epsilon}q is inside the dilation cone of Paneli,i+1\mathrm{Panel}_{i,i+1}. In this dilation cone, de=αiπd_{e}=\alpha_{i}\circ\pi. Thus, by definition of the Pansu derivative and the linearity of αi\alpha_{i},

P

​​D
de
|
p
(q)
=limϵ0de(pδϵq)de(p)ϵ=limϵ0αi(u+ϵv)1ϵ=αi(v)
.
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}|_{p}(q)=\lim_{\epsilon\to 0}\frac{d_{e}(p\delta_{\epsilon}q)-d_{e}(p)}{\epsilon}=\lim_{\epsilon\to 0}\frac{\alpha_{i}(u+\epsilon v)-1}{\epsilon}=\alpha_{i}(v).

5.4. Blow-ups of ded_{e} at non-smooth points

We now consider blow-ups of the function ded_{e} at points on the unit sphere which are not smooth, i.e., along the seams of the sphere.

5.4.1. Blow-ups near north and south poles

For each ii, define the cones

Ci=[0,+)𝗏i[0,+)𝗏i+1={𝗏iω0}{𝗏i+1ω0}C_{i}^{-}=-[0,+\infty)\mathsf{v}_{i}-[0,+\infty)\mathsf{v}_{i+1}=\{\mathsf{v}_{i}^{\omega}\leq 0\}\cap\{\mathsf{v}_{i+1}^{\omega}\geq 0\}

in 2\mathbb{R}^{2}. If vCiv\in C^{-}_{i}, then v=αi(v)\|v\|=-\alpha_{i}(v). Recall from [6] that the non-degenerate QijQ_{ij} containing (0,0)(0,0) are Qi+1,iQ_{i+1,i} for i=1,,2Ni=1,\dots,2N. For each ii we also define the dilation cones Ui:=δ(0,+)Paneli+1,i+U_{i}:=\delta_{(0,+\infty)}\mathrm{Panel}^{+}_{i+1,i}. Notice that

ui+1,i(r,s)\displaystyle u_{i+1,i}(r,s) =1μ(rσi+1+σi+2++σi1+sσi)\displaystyle=\frac{1}{\mu}(r\sigma_{i+1}+\sigma_{i+2}+\dots+\sigma_{i-1}+s\sigma_{i})
=1μ(rσi+1+σi+2++σi1+σi+σi+1+sσiσiσi+1)\displaystyle=\frac{1}{\mu}(r\sigma_{i+1}+\sigma_{i+2}+\dots+\sigma_{i-1}+\sigma_{i}+\sigma_{i+1}+s\sigma_{i}-\sigma_{i}-\sigma_{i+1})
=(1r)σi+1μvi+1(1s)σiμvi.\displaystyle=-(1-r)\frac{\|\sigma_{i+1}\|}{\mu}v_{i+1}-(1-s)\frac{\|\sigma_{i}\|}{\mu}v_{i}.

Therefore, π(Paneli+1,i+)=Qi+1,iCi\pi(\mathrm{Panel}^{+}_{i+1,i})=Q_{i+1,i}\subset C_{i}^{-}, and thus Uiπ1(Ci)U_{i}\subset\pi^{-1}(C^{-}_{i}).

Proposition 5.8 (Blow-ups at north and south poles).

Let pp be the north or south pole of the unit sphere B\partial B. Then, all blow-ups of ded_{e} at pp are:

  1. (1)

    For w2w\in\mathbb{R}^{2},

    f(v,t)=wwv;f(v,t)=\lVert w\rVert-\lVert w-v\rVert;
  2. (2)

    For C{,+}C\in\mathbb{R}\cup\{-\infty,+\infty\} and i{1,,2N}i\in\{1,\ldots,2N\}

    f(v,t)={αi(v)+c1ω(𝗏i,v)Cαi1(v)+c2ω(𝗏i,v)>C;f(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\leq C\\ \alpha_{i-1}(v)+c_{2}&\omega(\mathsf{v}_{i},v)>C\end{cases};

Proposition 5.8 gives a second proof of Theorem 4.1 in the case of polygonal sub-Finsler distances.

Proof.

Suppose pp is the north pole. A sufficiently small neighborhood Ω\Omega of pp is covered by the dilation cones UiU_{i}. Moreover, up to shrinking Ω\Omega, we can suppose UiΩ=π1(Ci)ΩU_{i}\cap\Omega=\pi^{-1}(C^{-}_{i})\cap\Omega.

From Proposition 3.7, we conclude that all blow-ups of UiU_{i} at pp are \mathbb{H}, \emptyset, left translations of π1(Ci)\pi^{-1}(C^{-}_{i}), and the half spaces π1({viω0})\pi^{-1}(\{v_{i}^{\omega}\leq 0\}) and π1({vi+1ω0})\pi^{-1}(\{v_{i+1}^{\omega}\geq 0\}).

Next, we see from (19), (20) and (21) that (r,s)(ui+1,i(r,s),ϕi+1,i(r,s))(r,s)\mapsto(u_{i+1,i}(r,s),\phi_{i+1,i}(r,s)) is well defined in a neighborhood of (1,1)(1,1) and the image is not tangent to [,][\mathbb{H},\mathbb{H}] at (1,1)(1,1). Notice that ui+1,i(1,1)=(0,0)u_{i+1,i}(1,1)=(0,0). It follows that the map ψi:(r,s,t)δt(ui+1,i(r,s),ϕi+1,i(r,s))\psi_{i}:(r,s,t)\mapsto\delta_{t}(u_{i+1,i}(r,s),\phi_{i+1,i}(r,s)) is a diffeomorphism near to (1,1,1)(1,1,1) and ψi(1,1,1)=p\psi_{i}(1,1,1)=p. Therefore, we can extend de|Uid_{e}|_{U_{i}} to a homogeneous smooth function fif_{i} defined in a neighborhood of pp by fi(q)=t(ψi1(q))f_{i}(q)=t(\psi_{i}^{-1}(q)). Using Proposition 5.6 and the smoothness of fif_{i}, we deduce that

P

​​D
fi
|
p
(v,t)
=αi(v)
.
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}f_{i}|_{p}(v,t)=\alpha_{i}(v).

We are now in the position to conclude the proof. On the one hand, if {pn}n\{p_{n}\}_{n\in\mathbb{N}}\subset\mathbb{H} and {ϵn}n(0,+)\{\epsilon_{n}\}_{n\in\mathbb{N}}\subset(0,+\infty) are sequences with pnpp_{n}\to p and ϵn0\epsilon_{n}\to 0, then, up to passing to a subsequence, we can assume that 𝙱𝚄(Ui,{pn}n,{ϵn}n)\mathtt{BU}(U_{i},\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) exist for each ii, by Theorem 3.1. Therefore, by Theorem 3.9, we obtain that 𝙱𝚄((,de),{pn}n,{ϵn}n)\mathtt{BU}((\mathbb{H},d_{e}),\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) is one of the functions listed in the statement.

On the other hand, if ff is one of the functions listed in the statement, then there are sequences {pn}n\{p_{n}\}_{n\in\mathbb{N}}\subset\mathbb{H} and {ϵn}n(0,+)\{\epsilon_{n}\}_{n\in\mathbb{N}}\subset(0,+\infty) with pnpp_{n}\to p and ϵn0\epsilon_{n}\to 0 so that the blow-ups 𝙱𝚄(Ui,{pn}n,{ϵn}n)\mathtt{BU}(U_{i},\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}) make the partition of \mathbb{H} given by ff and thus, by Proposition 3.7, we obtain (,f)=𝙱𝚄((,de),{pn}n,{ϵn}n)(\mathbb{H},f)=\mathtt{BU}((\mathbb{H},d_{e}),\{p_{n}\}_{n},\{\epsilon_{n}\}_{n}).

For the south pole the proof is the same. ∎

Proposition 5.9 (Blow-ups at the north star seam).

Let p=(uii(r,0),ϕii(r,0))p=(u_{ii}(r,0),\phi_{ii}(r,0)), r(0,1)r\in(0,1), be a ceiling point above the star of QQ in the degenerate panel Panelii+\mathrm{Panel}_{ii}^{+}. All the blow-ups of ded_{e} at pp are

f(v,t)={αi1(v)+c1ω(𝗏i,v)C((1r)αi1+rαi)(v)+c2ω(𝗏i,v)<Cf(v,t)=\begin{cases}\alpha_{i-1}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\geq C\\ ((1-r)\alpha_{i-1}+r\alpha_{i})(v)+c_{2}&\omega(\mathsf{v}_{i},v)<C\end{cases}

for C{,+}C\in\mathbb{R}\cup\{-\infty,+\infty\}.

Proof.

Notice that ui,i(r,0)=ui,i1(r,1)=ui+1,i(1,r)u_{i,i}(r,0)=u_{i,i-1}(r,1)=u_{i+1,i}(1,r), hence Qi,iQi,i1Qi+1,iQ_{i,i}\subset Q_{i,i-1}\cap Q_{i+1,i}. A sufficiently small neighborhood Ω\Omega of pp is covered by the two cones Ui1U_{i-1} and UiU_{i}. Up to shrinking Ω\Omega, we have Ui1Ω={𝗏iω0}ΩU_{i-1}\cap\Omega=\{\mathsf{v}_{i}^{\omega}\geq 0\}\cap\Omega and UiΩ={𝗏iω0}ΩU_{i}\cap\Omega=\{\mathsf{v}_{i}^{\omega}\leq 0\}\cap\Omega.

Thus, arguing like in the proof of Proposition 5.8, we can smoothly extend both de|Uid_{e}|_{U_{i}} and de|Ui+1d_{e}|_{U_{i+1}} to Ω\Omega and show that all the blow-ups of ded_{e} at pp are those listed in the statement. ∎

A similar analysis of points in star line segments in the basement of the unit sphere yields the following proposition.

Proposition 5.10 (Blow-ups at the south star seam).

Let p=(uii(r,0),ϕii(r,0))p=(u_{ii}(r,0),-\phi_{ii}(r,0)), r(0,1)r\in(0,1), be a basement point below the star of QQ in the degenerate panel Panelii\mathrm{Panel}_{ii}^{-}. All the blow-ups of ded_{e} at pp are

f(v,t)={αi(v)+c1ω(𝗏i,v)C((1r)αi+rαi1)(v)+c2ω(𝗏i,v)>C.f(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\leq C\\ ((1-r)\alpha_{i}+r\alpha_{i-1})(v)+c_{2}&\omega(\mathsf{v}_{i},v)>C\end{cases}.

for C{,+}C\in\mathbb{R}\cup\{-\infty,+\infty\}.

Proposition 5.11 (Blow-ups at the tips of the star seam).

Let pBp\in\partial B be such that π(p)=uii(1,0)\pi(p)=u_{ii}(1,0). Then ded_{e} is Pansu differentiable at pp.

Proof.

The point pp lies at the end of the star line segment and in the intersection of a third panel Paneli+1,i1±\mathrm{Panel}_{i+1,i-1}^{\pm}. Checking the (r,s)(r,s) coordinates of pp in the three panels, one sees that the three pieces of the blow-up function are all equal to αi1\alpha_{i-1}. Thus the Pansu derivative of d2d_{2} at pp exists. ∎

5.4.2. Blow-ups along wall seams

For each ii, define the cones

Ci+=[0,+)𝗏i+[0,+)𝗏i+1={𝗏iω0}{𝗏i+1ω0}C_{i}^{+}=[0,+\infty)\mathsf{v}_{i}+[0,+\infty)\mathsf{v}_{i+1}=\{\mathsf{v}_{i}^{\omega}\geq 0\}\cap\{\mathsf{v}_{i+1}^{\omega}\leq 0\}

in 2\mathbb{R}^{2}. If vCi+v\in C_{i}^{+}, then v=αi(v)\|v\|=\alpha_{i}(v). For each ii we also define the dilation cones Wi:=δ(0,+)Paneli,i+1+W_{i}:=\delta_{(0,+\infty)}\mathrm{Panel}^{+}_{i,i+1}, where Paneli,i+1+\mathrm{Panel}^{+}_{i,i+1} is the vertical wall of B\partial B containing the edge of QQ between viv_{i} and vi+1v_{i+1}. We recall that

ui,i+1(r,s)=rσirσi+sσi+1vi+sσi+1rσi+sσi+1vi+1u_{i,i+1}(r,s)=\frac{r\|\sigma_{i}\|}{r\|\sigma_{i}\|+s\|\sigma_{i+1}\|}v_{i}+\frac{s\|\sigma_{i+1}\|}{r\|\sigma_{i}\|+s\|\sigma_{i+1}\|}v_{i+1}

is a convex combination of viv_{i} and vi+1v_{i+1}.

The boundary of WiW_{i} is made up of a top and a bottom piece, each of which is smooth, which we denote by Wi+\partial W_{i}^{+} and Wi\partial W_{i}^{-}, respectively. There exists a function F^:Ci+\hat{F}:C_{i}^{+}\to\mathbb{R} whose graph is Wi+\partial W_{i}^{+} Indeed, Wi\partial W_{i} is parametrized by

Wi±={(ϵ((1λ)𝗏i+λ𝗏i+1),±ϵ22ω(𝗏i,𝗏i+1)(λλ2)):ϵ(0,),λ[0,1]}.\partial W_{i}^{\pm}=\{(\epsilon((1-\lambda)\mathsf{v}_{i}+\lambda\mathsf{v}_{i+1}),\pm\frac{\epsilon^{2}}{2}\omega(\mathsf{v}_{i},\mathsf{v}_{i+1})(\lambda-\lambda^{2})):\epsilon\in(0,\infty),\lambda\in[0,1]\}.

Using this parametrization, we solve for the height function,

F^i(v)=ω(𝗏i,v)ω(v,𝗏i+1)2ω(𝗏i,𝗏i+1).\hat{F}_{i}(v)=\frac{\omega(\mathsf{v}_{i},v)\omega(v,\mathsf{v}_{i+1})}{2\omega(\mathsf{v}_{i},\mathsf{v}_{i+1})}.

Thus, Wi={Fi0}W_{i}=\{F_{i}\leq 0\}, where Fi(v,t)=|t|F^i(v)F_{i}(v,t)=|t|-\hat{F}_{i}(v), which is smooth except in the {t=0}\{t=0\} plane. Notice that

(25)

P

​​D
Fi
|
(w,s)
(v,t)
={ω(w,v)2ω(𝗏i,v)ω(w,𝗏i+1)+ω(𝗏i,w)ω(v,𝗏i+1)2ω(𝗏i,𝗏i+1) if s>0,ω(w,v)2ω(𝗏i,v)ω(w,𝗏i+1)+ω(𝗏i,w)ω(v,𝗏i+1)2ω(𝗏i,𝗏i+1) if s<0.
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{i}|_{(w,s)}(v,t)=\begin{cases}\frac{\omega(w,v)}{2}-\frac{\omega(\mathsf{v}_{i},v)\omega(w,\mathsf{v}_{i+1})+\omega(\mathsf{v}_{i},w)\omega(v,\mathsf{v}_{i+1})}{2\omega(\mathsf{v}_{i},\mathsf{v}_{i+1})}&\text{ if }s>0,\\ -\frac{\omega(w,v)}{2}-\frac{\omega(\mathsf{v}_{i},v)\omega(w,\mathsf{v}_{i+1})+\omega(\mathsf{v}_{i},w)\omega(v,\mathsf{v}_{i+1})}{2\omega(\mathsf{v}_{i},\mathsf{v}_{i+1})}&\text{ if }s<0.\end{cases}

If w=λ𝗏i+(1λ)𝗏i+1w=\lambda\mathsf{v}_{i}+(1-\lambda)\mathsf{v}_{i+1}, then

(26)

P

​​D
Fi
|
(w,s)
(v,t)
={(1λ)ω(𝗏i+1,v) if s>0,λω(v,𝗏i) if s<0.
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}F_{i}|_{(w,s)}(v,t)=\begin{cases}(1-\lambda)\omega(\mathsf{v}_{i+1},v)&\text{ if }s>0,\\ \lambda\omega(v,\mathsf{v}_{i})&\text{ if }s<0.\end{cases}
Proposition 5.12 (Blow-ups along wall seams: ceiling).

Let pp be a point on B\partial B which lies on the seam between the vertical side Paneli,i+1\mathrm{Panel}_{i,i+1} and the ceiling such that π(p)=ui,i+1(r,s)\pi(p)=u_{i,i+1}(r,s) with r,s(0,1]r,s\in(0,1], one of them equal to 1. Then all the blow-ups of ded_{e} at pp are

f(v,t)={αi(v)+c1ω(𝗏i+1,v)C((1s)αi+sαi+1)(v)+c2ω(𝗏i+1,v)>C,f(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i+1},v)\leq C\\ ((1-s)\alpha_{i}+s\alpha_{i+1})(v)+c_{2}&\omega(\mathsf{v}_{i+1},v)>C\end{cases},

for C{+,}C\in\mathbb{R}\cup\{+\infty,-\infty\}.

Proof.

We consider three cases. First, if r=1r=1 and s<1s<1, then π(p)=ui,i+1(1,s)=ui1,i+1(0,s)Qi1,i+1\pi(p)=u_{i,i+1}(1,s)=u_{i-1,i+1}(0,s)\in Q_{i-1,i+1}. Therefore, a neighborhood Ω\Omega of pp is decomposed into two regions, ΩWi=Ω{Fi0}\Omega\cap W_{i}=\Omega\cap\{F_{i}\leq 0\} and Ωδ(0,+)Paneli1,i+1+=Ω{Fi0}\Omega\cap\delta_{(0,+\infty)}\mathrm{Panel}^{+}_{i-1,i+1}=\Omega\cap\{F_{i}\geq 0\}. Therefore, Propositions 3.6 and 3.7 with (26) show what are all the blow-ups of this decomposition, while Propositions 5.6 and 5.7 give us the Pansu differentials of ded_{e} near to pp, so that we can conclude using Theorem 3.9.

Second, if r<1r<1 and s=1s=1, then π(p)=ui,i+1(r,s)=ui,i+2(r,0)Qi,i+2\pi(p)=u_{i,i+1}(r,s)=u_{i,i+2}(r,0)\in Q_{i,i+2}. Then we can proceed like before.

Third, when r=s=1r=s=1, then the point pp lies in the boundary of four regions: WiW_{i}, δ(0,+)Paneli1,i+1+\delta_{(0,+\infty)}\mathrm{Panel}^{+}_{i-1,i+1}, δ(0,+)Paneli,i+2+\delta_{(0,+\infty)}\mathrm{Panel}^{+}_{i,i+2} and δ(0,+)Paneli1,i+2+\delta_{(0,+\infty)}\mathrm{Panel}^{+}_{i-1,i+2}. However, the function ded_{e} is CH1C^{1}_{H} on the union of the latter three and

P

​​D
de
\text{\scalebox{-1.0}[1.0]{$\mathrm{P}$}\!\!$\mathrm{D}$}d_{e}
has a continuous extension to pp. It follows that the blow-ups of ded_{e} at pp are again the ones listed above. This completes the proof. ∎

A similar result holds for the basement.

Proposition 5.13 (Blow-ups along wall seams: basement).

Let pp be a point on B\partial B which lies on the seam between the vertical side Paneli,i+1\mathrm{Panel}_{i,i+1} and the basement such that π(p)=ui,i+1(r,s)\pi(p)=u_{i,i+1}(r,s) with r,s(0,1]r,s\in(0,1], one of them equal to 1. Then all the blow-ups of ded_{e} at pp are

f(v,t)={αi(v)+c1ω(𝗏i,v)C((1r)αi+rαi1)(v)+c2ω(𝗏i,v)>C,f(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\leq C\\ ((1-r)\alpha_{i}+r\alpha_{i-1})(v)+c_{2}&\omega(\mathsf{v}_{i},v)>C\end{cases},

for C{+,}C\in\mathbb{R}\cup\{+\infty,-\infty\}.

Proposition 5.14 (Blow-ups along wall seams: vertices).

Let pp be the vertex 𝗏i\mathsf{v}_{i} on the unit sphere. Then all the blow-ups of ded_{e} at pp are

f(v,t)={αi(v)+c1ω(𝗏i,v)Cαi1(v)+c2ω(𝗏i,v)<C,f(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\geq C\\ \alpha_{i-1}(v)+c_{2}&\omega(\mathsf{v}_{i},v)<C\end{cases},

for C{+,}C\in\mathbb{R}\cup\{+\infty,-\infty\}.

Proof.

The point pp belongs to four regions: the two wall cones WiW_{i} and Wi1W_{i-1}, and the cones Pi+:=δ(0,+)Paneli1,i+1+P_{i}^{+}:=\delta_{(0,+\infty)}\mathrm{Panel}^{+}_{i-1,i+1} and Pi:=δ(0,+)Paneli1,i+1P_{i}^{-}:=\delta_{(0,+\infty)}\mathrm{Panel}^{-}_{i-1,i+1} Using the formulas for FiF_{i} written above, one readily sees that

Klimϵ0δ1/ϵ((𝗏i,0)11Pi+)\displaystyle\operatorname*{K-lim}_{\epsilon\to 0}\delta_{1/\epsilon}((\mathsf{v}_{i},0)^{-1}{-1}P_{i}^{+}) ={(v,t):ω(𝗏i,v)0}{(v,t):tF^i(v)0},\displaystyle=\{(v,t):\omega(\mathsf{v}_{i},v)\geq 0\}\cap\{(v,t):t-\hat{F}_{i}(v)\geq 0\},
Klimϵ0δ1/ϵ((𝗏i,0)11Wi)\displaystyle\operatorname*{K-lim}_{\epsilon\to 0}\delta_{1/\epsilon}((\mathsf{v}_{i},0)^{-1}{-1}W_{i}) ={(v,t):ω(𝗏i,v)0}{(v,t):tF^i(v)0},\displaystyle=\{(v,t):\omega(\mathsf{v}_{i},v)\geq 0\}\cap\{(v,t):t-\hat{F}_{i}(v)\leq 0\},

The union of these two limit cones is the half-space {(v,t):ω(𝗏i,v)0}\{(v,t):\omega(\mathsf{v}_{i},v)\geq 0\}. Similarly, PiP_{i}^{-} and Wi1W_{i-1} blow-up to {(v,t):ω(𝗏i,v)0}\{(v,t):\omega(\mathsf{v}_{i},v)\leq 0\}.

Meanwhile, the function ded_{e} blows up to αi\alpha_{i} on both in Pi+P_{i}^{+} and WiW_{i}, while it blows up to αi1\alpha_{i-1} on both PiP_{i}^{-} and Wi1W_{i-1}. We then conclude. ∎

6. Dynamics of the action of \mathbb{H} on the boundary

One of the main motivations for studying the boundary of a metric space is to then examine how the group of isometries acts on the boundary. Ideally this action on the boundary is simpler than the action on the space itself, and one can hope to glean information about the space or the group through this action. In any Lie group with a left-invariant metric, the group acts isometrically on itself via left translation. In this section, we explore how \mathbb{H} with a polygonal sub-Finsler metric acts on its horofunction boundary and its reduced horofunction boundary, generalizing results on finitely generated nilpotent groups by Walsh and Bader-Finkelshtein [24, 1]. Given that our polygonal sub-Finsler metrics are the asymptotic cones of the discrete word metrics, the fact that the results generalize is not overly surprising.

6.1. Action of the group on the boundary

Let dd be any left-invariant homogeneous metric on \mathbb{H}. To understand how the group acts on the boundary, it suffices to understand how the group acts on sequences. Suppose {qn}n\{q_{n}\}_{n} is a sequence in \mathbb{H} which converges to a horofunction fh(,d)f\in\partial_{h}(\mathbb{H},d). By definition f(x)=limnd(qn,x)d(qn,e)f(x)=\lim_{n\to\infty}d(q_{n},x)-d(q_{n},e). For a group element gg\in\mathbb{H}, the image g.f(x)g.f(x) is the limit of the translated sequence {gqn}n\{gq_{n}\}_{n}. We have

g.f(x)\displaystyle g.f(x) =limnd(gqn,x)d(gqn,e)\displaystyle=\lim_{n\to\infty}d(gq_{n},x)-d(gq_{n},e)
=limnd(qn,g1x)d(qn,e)d(qn,g1)+d(qn,e)\displaystyle=\lim_{n\to\infty}d(q_{n},g^{-1}x)-d(q_{n},e)-d(q_{n},g^{-1})+d(q_{n},e)
=f(g1x)f(g1).\displaystyle=f(g^{-1}x)-f(g^{-1}).

In Lemma 2.3, we observed how horofunctions are related to Pansu derivatives and blow-ups of the distance function at points on the unit sphere. In particular, we have shown any horofunction ff in the boundary can be realized as a limit

f(x)=limnde(pnδϵnx)de(pn)ϵn,f(x)=\lim_{n\to\infty}\frac{d_{e}(p_{n}\delta_{\epsilon_{n}}x)-d_{e}(p_{n})}{\epsilon_{n}},

where pnpBp_{n}\to p\in\partial B and ϵn0\epsilon_{n}\to 0. The following lemma shows that g.fg.f similarly is a directional derivative of ded_{e} at the same point pp.

Lemma 6.1.

Suppose fh(,d)f\in\partial_{h}(\mathbb{H},d) is a blow-up of ded_{e} at a point pp on the unit sphere B\partial B. Then for any gg\in\mathbb{H}, the boundary point g.fg.f is also a blow-up of ded_{e} at pp.

Proof.

Let {pn}n\{p_{n}\}_{n}, pnpp_{n}\to p, and {ϵn}n\{\epsilon_{n}\}_{n}, ϵn0\epsilon_{n}\to 0, be such that

f(x)=limnde(pnδϵnx)de(pn)ϵn.f(x)=\lim_{n\to\infty}\frac{d_{e}(p_{n}\delta_{\epsilon_{n}}x)-d_{e}(p_{n})}{\epsilon_{n}}.

The corresponding sequence in \mathbb{H} which converges to ff is {qn}n={δ1/ϵnpn1}n\{q_{n}\}_{n}=\{\delta_{1/\epsilon_{n}}p_{n}^{-1}\}_{n}. Since δϵnqn=pn1p1\delta_{\epsilon_{n}}q_{n}=p_{n}^{-1}\to p^{-1}, we say that qnq_{n} converges in direction to p1p^{-1}. This observation implies that ff is a blow-up of ded_{e} along a sequence pnpp_{n}\to p if and only if the sequence {qn}n\{q_{n}\}_{n} which converges to ff converges in direction to p1p^{-1}. If we translate {qn}n\{q_{n}\}_{n} by an element gg\in\mathbb{H}, we observe that

δϵn(gqn)=(δϵng)(δϵnqn)p1.\delta_{\epsilon_{n}}(gq_{n})=(\delta_{\epsilon_{n}}g)(\delta_{\epsilon_{n}}q_{n})\to p^{-1}.

Thus gqngq_{n} also converges in direction to p1p^{-1}, and g.fg.f is the blow-up of ded_{e} at pp along the sequence {pnδϵng1}n\{p_{n}\delta_{\epsilon_{n}}g^{-1}\}_{n} with the same {ϵn}n\{\epsilon_{n}\}_{n}. ∎

Remark 6.2.

We recall from Proposition 3.8 that ded_{e} is strictly Pansu differentiable at a point pp, then the blow-up of ded_{e} along any sequences {pn}n\{p_{n}\}_{n} and {ϵn}n\{\epsilon_{n}\}_{n} satisfying pnpp_{n}\to p and ϵn0\epsilon_{n}\to 0 is equal to the Pansu derivative of ded_{e} at pp. Along with the previous lemma, this implies that if qnfh(,d)q_{n}\to f\in\partial_{h}(\mathbb{H},d) and qnq_{n} converges in direction to a point pp where the distance function ded_{e} is strictly Pansu differentiable, then g.f=fg.f=f for any gg\in\mathbb{H}.

6.2. Busemann functions

Recall that Busemann functions are points of the horofunction boundary which can be realized as limits of geodesic rays. In Corollary 5.5, we observe that the set of Busemann functions in the boundary of a polygonal sub-Finsler metric on \mathbb{H} is homeomorphic to a circle. Indeed, Busemann functions come in two flavors depending on whether they are the blow-ups of vertical wall points or vertices of the unit sphere:

f(q)=f(v,t)=αi(v) or f(q)=f(v,t)={αi(v)+c1ω(𝗏i,v)Cαi1(v)+c2ω(𝗏i,v)<C,f(q)=f(v,t)=\alpha_{i}(v)\quad\text{ or }\quad f(q)=f(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\geq C\\ \alpha_{i-1}(v)+c_{2}&\omega(\mathsf{v}_{i},v)<C\end{cases},

where i{1,,2N}i\in\{1,\ldots,2N\}, CC\in\mathbb{R}, and c1,c2c_{1},c_{2} are functions of CC, determined uniquely by the criteria f(e)=0f(e)=0 and ff is continuous.

In [24], Walsh proves that for any finitely generated nilpotent group, there is a one-to-one correspondence between finite orbits of Busemann functions under the action of the group and facets of a polyhedron defined by the generators of the group. The following proposition generalizes this result to the real Heisenberg group for any polygonal sub-Finsler metric.

Proposition 6.3.

In the boundary of a polygonal sub-Finsler metric on \mathbb{H}, there is a one-to-one correspondence between finite orbits of Busemann functions and edges of the metric-inducing polygon QQ.

Proof.

By Remark 6.2 and also by direct calculation, the action of the group on horofunctions of the form (v,t)αi(v)(v,t)\mapsto\alpha_{i}(v) is trivial. Since the αi\alpha_{i} are the blow-ups of ded_{e} on vertical walls of the unit sphere, we get a correspondence between the facets of QQ and finite orbits of the action.

It remains to show that no other Busemann functions are fixed globally by the action of the group. For each vertex 𝗏i\mathsf{v}_{i} we have a family of blow-ups, in this case Busemann functions,

i={f(v,t)={αi(v)+c1ω(𝗏i,v)Cαi1(v)+c2ω(𝗏i,v)<C:C}.\mathcal{F}_{i}=\left\{f(v,t)=\begin{cases}\alpha_{i}(v)+c_{1}&\omega(\mathsf{v}_{i},v)\geq C\\ \alpha_{i-1}(v)+c_{2}&\omega(\mathsf{v}_{i},v)<C\end{cases}:C\in\mathbb{R}\right\}.

A direct calculation shows that if g=(w,s)g=(w,s), fif\in\mathcal{F}_{i}, and ω(𝗏i,w)0\omega(\mathsf{v}_{i},w)\neq 0, then g.fig.f\in\mathcal{F}_{i}, but g.ffg.f\neq f. ∎

6.3. Trivial action on reduced horofunction boundary

When defining the horofunction boundary of a metric space, we defined the maps ι:X𝒞(X)\iota:X\hookrightarrow\mathscr{C}(X) and ι^:X𝒞(X)/\hat{\iota}:X\hookrightarrow\mathscr{C}(X)/\mathbb{R}. To define the reduced horofunction boundary we consider the image of h(X,d)\partial_{h}(X,d) in 𝒞(X)/𝒞b(X)\mathscr{C}(X)/\mathscr{C}_{b}(X), where 𝒞b(X)\mathscr{C}_{b}(X) is the space of all continuous bounded functions. It is worth noting that the reduced horofunction is not necessarily Hausdorff, but as we show below, it has value in its strong relationship with the action of the group on h(X,d)\partial_{h}(X,d).

In [1], Bader–Finkelshtein show that the for any finitely generated abelian group and discrete Heisenberg group with any finite generating set, the action of the group on its reduced horofunction boundary is trivial. They further conjecture that this result should hold for any finitely generated nilpotent group. We are able to extend this result to the real Heisenberg group with a polygonal sub-Finsler metric.

Proposition 6.4.

Let dd be a polygonal sub-Finsler metric on \mathbb{H}. Then the reduced horofunction boundary is in bijection with the quotient of h(,d)\partial_{h}(\mathbb{H},d) by the action of the group. That is

hr(,d)h(,d)/,\partial_{h}^{r}(\mathbb{H},d)\leftrightarrow\partial_{h}(\mathbb{H},d)/\mathbb{H},

and so \mathbb{H} acts trivially on its reduced horofunction boundary.

Proof.

To prove this proposition, it will suffice to look at each of the families of functions described in the Theorem 5.2.

We start by considering the three smooth families of horofunctions, which compose a circle in the boundary. These boundary points are all Pansu derivatives, and hence are linear. It is clear that two linear functions stay bounded distance from one another if and only if they are identical, and so each Pansu derivative remains distinct in the reduced horofunction boundary. By the definition of action on the boundary, it is clear that if ff is linear, then g.f=fg.f=f for all gg\in\mathbb{H}, and so the action on these points in hr(,d)\partial_{h}^{r}(\mathbb{H},d) is trivial.

Next we consider the piecewise-linear horofunctions coming from the blow-ups of non-smooth points. Any (nontrivially) piecewise linear function cannot have bounded difference from a linear function, and so they cannot be equivalent in the reduced horofunction boundary to the smooth families mentioned above. Our goal is to show that two horofunctions f1f_{1} and f2f_{2} differ by a bounded function if and only if f1f_{1} and f2f_{2} belong to the same orbit. Let f1f_{1} and f2f_{2} be distinct functions coming from the same family of functions in Theorem 5.2.

Case 1: Suppose that for j=1,2j=1,2, we have fj=wjwjvf_{j}=\lVert w_{j}\rVert-\lVert w_{j}-v\rVert, w1w2w_{1}\neq w_{2}. Then

|f2(v,t)f1(v,t)|\displaystyle|f_{2}(v,t)-f_{1}(v,t)| =|w2w2v(w1w1v)|\displaystyle=|\lVert w_{2}\rVert-\lVert w_{2}-v\rVert-(\lVert w_{1}\rVert-\lVert w_{1}-v\rVert)|
|w2w1|+|w1vw2v|\displaystyle\leq|\lVert w_{2}\rVert-\lVert w_{1}\rVert|+|\lVert w_{1}-v\rVert-\lVert w_{2}-v\rVert|
2w2w1,\displaystyle\leq 2\lVert w_{2}-w_{1}\rVert,

and so f1f_{1} and f2f_{2} are identified in the reduced boundary. It remains to show that they lie in the same orbit. Let g=(w2w1,0)g=(w_{2}-w_{1},0). Then

g.f1(v,t)\displaystyle g.f_{1}(v,t) =w1w1(v(w2w1))(w1w1(w2w1))\displaystyle=\lVert w_{1}\rVert-\lVert w_{1}-(v-(w_{2}-w_{1}))\rVert-(\lVert w_{1}\rVert-\lVert w_{1}-(w_{2}-w_{1})\rVert)
=w2w2v=f2(v,t).\displaystyle=\lVert w_{2}\rVert-\lVert w_{2}-v\rVert=f_{2}(v,t).

This calculation also shows that for any gg\in\mathbb{H}, g.f1g.f_{1} will lie in this same family of blow-ups.

Case 2: The remaining families of functions are the images of (0,1]×(0,1]\times\mathbb{R} under the maps ψi\psi_{i}^{\vee}, ψi\psi_{i}^{\wedge}, ξi\xi_{i}^{\vee}, and ξi\xi_{i}^{\wedge}, i{1,,2N}i\in\{1,\ldots,2N\}. It is clear that no function from these families can have bounded difference with a function from Case 1. Indeed, the norm-like functions of Case 1 are piecewise linear, where 2N2N distinct functions are defined on 2N2N regions, for N>1N>1. Meanwhile, the images of ψi\psi_{i}^{\vee}, ψi\psi_{i}^{\wedge}, ξi\xi_{i}^{\vee}, and ξi\xi_{i}^{\wedge} are defined by two functions defined on two halfspaces. There is, therefore, an unbounded region in the plane where the two functions have unbounded difference.

Consider two functions f1f_{1} and f2f_{2} from these families, where fjf_{j}, j=1,2,j=1,2, is the image of (sj,aj)(s_{j},a_{j}), sj(0,1]s_{j}\in(0,1], aja_{j}\in\mathbb{R}, under a map in {ψij,ψij,ξij,ξij}\{\psi_{i_{j}}^{\vee},\psi_{i_{j}}^{\wedge},\xi_{i_{j}}^{\vee},\xi_{i_{j}}^{\wedge}\}, for index ij{1,,2N}i_{j}\in\{1,\ldots,2N\}. We omit the cases where sj=0s_{j}=0 or aj{,}a_{j}\in\{-\infty,\infty\}, as these cases result in linear functions, already discussed above. We claim that f1f_{1} and f2f_{2} have bounded difference if and only 1) s1=s2s_{1}=s_{2}; 2) i1=i2i_{1}=i_{2}; and 3) they both are the image under the same map. Indeed, if any of these three conditions is not met, a direct inspection of the functions ψi\psi_{i}^{\vee}, ψi\psi_{i}^{\wedge}, ξi\xi_{i}^{\vee}, and ξi\xi_{i}^{\wedge} makes clear that there is an unbounded region on which f1f_{1} and f2f_{2} are defined as distinct linear functions and have unbounded difference. On the other hand, if the three conditions are met, we must show that f1f_{1} and f2f_{2} have bounded difference. For j=1,2j=1,2, let fjf_{j} be linear on the two regions Uj={(v,t):ω(𝗏i,v)Cj}U_{j}=\{(v,t):\omega(\mathsf{v}_{i},v)\geq C_{j}\} and Lj={(v,t):ω(𝗏i,v)Cj}L_{j}=\{(v,t):\omega(\mathsf{v}_{i},v)\geq C_{j}\}. To analyze the difference between f1f_{1} and f2f_{2}, we assume C1>C2C_{1}>C_{2} and consider three regions:

Ω1\displaystyle\Omega_{1} =U1U2={ω(𝗏i,v)C1}\displaystyle=U_{1}\cap U_{2}=\{\omega(\mathsf{v}_{i},v)\geq C_{1}\} Ω3=L1L2={ω(𝗏i,v)C2}\displaystyle\Omega_{3}=L_{1}\cap L_{2}=\{\omega(\mathsf{v}_{i},v)\leq C_{2}\}
Ω2\displaystyle\Omega_{2} =L1U2={C2<ω(𝗏i,v)<C1}\displaystyle=L_{1}\cap U_{2}=\{C_{2}<\omega(\mathsf{v}_{i},v)<C_{1}\}

Since s1=s2s_{1}=s_{2}, i1=i2i_{1}=i_{2}, and f1f_{1}, f2f_{2} are both images under the same map, (f1f2)|ω1(f_{1}-f_{2})|_{\omega_{1}} and (f1f2)|ω3(f_{1}-f_{2})|_{\omega_{3}} are constant. Meanwhile f1f_{1} and f2f_{2} are distinct on the unbounded strip Ω2\Omega_{2}. Since f1f_{1} and f2f_{2} are continuous, the level sets of the fjf_{j} in Ω2\Omega_{2} must be transverse (not parallel) to the strip Ω2\Omega_{2}, guaranteeing that the functions have bounded difference.

Finally, by choosing an element gg\in\mathbb{H} such that ω(𝗏i,π(g))=C1C2\omega(\mathsf{v}_{i},\pi(g))=C_{1}-C_{2}, one can confirm that f1f_{1} and f2f_{2} are in the same orbit. ∎

Refer to caption
Figure 11. The standard and reduced horofunction boundaries for a hexagonal sub-Finsler metric.

References

  • [1] Uri Bader and Vladimir Finkelshtein. On the horofunction boundary of discrete Heisenberg group. Geometriae Dedicata, pages 1–15, 2020.
  • [2] Martin R. Bridson and André Haefliger. Metric spaces of non-positive curvature, volume 319 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999.
  • [3] Herbert Busemann. The isoperimetric problem in the Minkowski plane. Amer. J. Math., 69:863–871, 1947.
  • [4] Corina Ciobotaru, Linus Kramer, and Petra Schwer. Polyhedral compactifications, i. arXiv preprint arXiv:2002.12422, 2020.
  • [5] Moon Duchin and Nathan Fisher. Stars at infinity in Teichmüller space. arXiv preprint arXiv:2004.04231, 2020.
  • [6] Moon Duchin and Christopher Mooney. Fine asymptotic geometry in the Heisenberg group. Indiana Univ. Math. J., 63(3):885–916, 2014.
  • [7] G. B. Folland and Elias M. Stein. Hardy spaces on homogeneous groups, volume 28 of Mathematical Notes. Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1982.
  • [8] V. Gershkovich and A. Vershik. Nonholonomic manifolds and nilpotent analysis. J. Geom. Phys., 5(3):407–452, 1988.
  • [9] Lizhen Ji and Anna-Sofie Schilling. Polyhedral horofunction compactification as polyhedral ball. arXiv preprint arXiv:1607.00564, 2016.
  • [10] Antoine Julia, Sebastiano Nicolussi Golo, and Davide Vittone. Area of intrinsic graphs and coarea formula in Carnot Groups. arXiv e-prints, page arXiv:2004.02520, April 2020.
  • [11] Anders Karlsson, Volker Metz, and Gennady A Noskov. Horoballs in simplices and Minkowski spaces. International journal of mathematics and mathematical sciences, 2006, 2006.
  • [12] Tom Klein and Andrew Nicas. The horofunction boundary of the Heisenberg group. Pacific journal of mathematics, 242(2):299–310, 2009.
  • [13] Tom Klein and Andrew Nicas. The horofunction boundary of the Heisenberg group: the Carnot-Carathéodory metric. Conformal Geometry and Dynamics of the American Mathematical Society, 14(15):269–295, 2010.
  • [14] Enrico Le Donne. A primer on Carnot groups: homogenous groups, Carnot-Carathéodory spaces, and regularity of their isometries. Anal. Geom. Metr. Spaces, 5(1):116–137, 2017.
  • [15] Enrico Le Donne and Sebastiano Nicolussi Golo. Regularity properties of spheres in homogeneous groups. Trans. Amer. Math. Soc., 370(3):2057–2084, 2018.
  • [16] Enrico Le Donne, Sebastiano Nicolussi Golo, and Andrea Sambusetti. Asymptotic behavior of the Riemannian Heisenberg group and its horoboundary. Ann. Mat. Pura Appl. (4), 196(4):1251–1272, 2017.
  • [17] Enrico Le Donne and Séverine Rigot. Besicovitch covering property on graded groups and applications to measure differentiation. J. Reine Angew. Math., 750:241–297, 2019.
  • [18] John Mitchell. On Carnot-Carathéodory metrics. J. Differential Geom., 21(1):35–45, 1985.
  • [19] John William Mitchell. A local study of Carnot-Caratheodory metrics. ProQuest LLC, Ann Arbor, MI, 1982. Thesis (Ph.D.)–State University of New York at Stony Brook.
  • [20] Alexander Nagel, Elias M. Stein, and Stephen Wainger. Balls and metrics defined by vector fields. I. Basic properties. Acta Math., 155(1-2):103–147, 1985.
  • [21] Pierre Pansu. Croissance des boules et des géodésiques fermées dans les nilvariétés. Ergodic Theory Dynam. Systems, 3(3):415–445, 1983.
  • [22] Pierre Pansu. Métriques de Carnot-Carathéodory et quasiisométries des espaces symétriques de rang un. Ann. of Math. (2), 129(1):1–60, 1989.
  • [23] Cormac Walsh. The horofunction boundary of finite-dimensional normed spaces. Math. Proc. Cambridge Philos. Soc., 142(3):497–507, 2007.
  • [24] Cormac Walsh. The action of a nilpotent group on its horofunction boundary has finite orbits. Groups Geom. Dyn., 5(1):189–206, 2011.
  • [25] Kazimierz Zarankiewicz. Sur les points de division dans les ensembles connexes. Uniwersytet, Seminarjum Matematyczne, 1927.