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

On Boundaries of -neighbourhoods
of Planar Sets, Part I: Singularities

Jeroen S.W. Lamb Jeroen S.W. Lamb
Department of Mathematics
Imperial College London
180 Queen's Gate, London SW7 2AZ, United Kingdom
[email protected]
Martin Rasmussen Martin Rasmussen
Department of Mathematics
Imperial College London
180 Queen's Gate, London SW7 2AZ, United Kingdom
[email protected]
 and  Kalle Timperi Kalle Timperi
Department of Mathematics
Imperial College London
180 Queen's Gate, London SW7 2AZ, United Kingdom
[email protected]
Abstract.

We study geometric and topological properties of singularities on the boundaries of ε\varepsilon-neighbourhoods Eε={x2:dist(x,E)ε}E_{\varepsilon}=\{x\in\mathbb{R}^{2}\,:\,\textrm{dist}(x,E)\leq\varepsilon\} of planar sets E2E\subset\mathbb{R}^{2}. We develop a novel technique for analysing the boundary and obtain, for a compact set EE and ε>0\varepsilon>0, a classification of singularities (i.e. non-smooth points) on Eε\partial E_{\varepsilon} into eight categories. We also show that the set of singularities is either countable or the disjoint union of a countable set and a closed, totally disconnected, nowhere dense set.

2020 Mathematics Subject Classification:
51F30; 57K20, 54C50, 51M15, 58C06

1. Introduction

1.1. Motivation

For a given set EdE\subset\mathbb{R}^{d} and radius ε>0\varepsilon>0, the (closed) ε\varepsilon-neighbourhood of EE is the set

(1.1) Eε:=Bε(E)¯:=xEBε(x)¯,E_{\varepsilon}:=\overline{B_{\varepsilon}(E)}:=\overline{\bigcup_{x\in E}B_{\varepsilon}(x)},

where the overline denotes closure and Bε()B_{\varepsilon}(\cdot) is an open ball of radius ε\varepsilon in the Euclidean metric. The sets EεE_{\varepsilon} are also known in the literature as tubular neighbourhoods [15], collars [24] or parallel sets [27, 33]. The boundary Eε\partial E_{\varepsilon} is a subset of the set E<ε:=(xEBε(x))\partial E_{<\varepsilon}:=\partial\left(\bigcup_{x\in E}B_{\varepsilon}(x)\right), which is sometimes referred to as the ε\varepsilon-boundary [14] or ε\varepsilon-level set [23] of EE.

The central question addressed in this paper concerns the geometric and topological properties of such sets EεE_{\varepsilon}, with a focus on properties of its boundary Eε\partial E_{\varepsilon}. This is not only a very natural and fundamental question in (Euclidean) geometry, but it is also relevant in specific settings where ε\varepsilon-neighbourhoods naturally arise. For instance, we are motivated by the classification and bifurcation of minimal invariant sets in random dynamical systems with bounded noise [21], but ε\varepsilon-neighbourhoods also naturally feature for instance in control theory [9].

Notwithstanding significant theoretical progress on the properties of ε\varepsilon-neighbourhoods during the last decades, the geometric classification of possible boundaries Eε\partial E_{\varepsilon} has remained open, even in dimension two.

1.2. Main results

Refer to caption
Figure 1. Schematic illustration of Theorem 1. The grey area represents the ε\varepsilon-neighbourhood EεE_{\varepsilon}, the white area the complement 2Eε\mathbb{R}^{2}\setminus E_{\varepsilon}. Every boundary point xEεx\in\partial E_{\varepsilon} either is a smooth point or belongs to exactly one of eight categories of singularities. At a wedge (S1) the one-sided tangents form an angle 0<θ<π0<\theta<\pi. A sharp singularity (S2) and a sharp-sharp singularity (S3) can be thought of as extremal cases of a wedge, with θ=0\theta=0. A shallow singularity (S4) and a shallow-shallow singularity (S5) have a well-defined tangent, but they are accumulation points (from one or two directions, respectively) of sequences of increasingly obtuse wedges (black dots). A chain singularity (S6), a chain-chain singularity (S7) and a sharp-chain singularity (S8) share the geometric property of being accumulation points of sequences of increasingly acute wedges (black dots). This turns out to be equivalent (see Proposition 4.5) to the topological property of being the limit with respect to Hausdorff distance (see Definition 3.2) of a sequence of disjoint connected components of the complement EεcE_{\varepsilon}^{c}. See also Figure 9.

Our main achievement is the development of a novel technique for analysing geometric properties of the boundary Eε\partial E_{\varepsilon}, enabling a local representation for the boundary around every boundary point xEεx\in\partial E_{\varepsilon} via graphs of Lipschitz continuous functions. We employ this representation to obtain a classification of points on the boundary of ε\varepsilon-neighbourhoods of compact planar sets.

Our first main result establishes that for any compact set E2E\in\mathbb{R}^{2} and ε>0\varepsilon>0, each boundary point xEεx\in\partial E_{\varepsilon} is either a smooth point (in the sense that, in a neighbourhood of xx, Eε\partial E_{\varepsilon} is a C1C^{1}-curve) or falls into exactly one of eight distinct categories of singularities.

Theorem 1.

Let E2E\subset\mathbb{R}^{2} be compact, ε>0\varepsilon>0, and let xEεx\in\partial E_{\varepsilon} be a boundary point of EεE_{\varepsilon} that is not smooth. Then xx belongs to precisely one of the following eight categories:

(S1) wedge, (S5) shallow-shallow singularity,
(S2) sharp singularity, (S6) chain singularity,
(S3) sharp-sharp singularity, (S7) chain-chain singularity,
(S4) shallow singularity, (S8) sharp-chain singularity.

Definition 4.1 contains a rigorous definition of these categories, but for indicative sketches of the singularity types, see Figure 1.

The proof of Theorem 1 is based on the construction of a local boundary representation, given in Proposition 3.5, that allows us to treat small parts of the boundary Eε\partial E_{\varepsilon} as finite unions of graphs of continuous functions. This representation in turn relies on a local contribution property, Proposition 3.1, which states that the geometry of the boundary near each boundary point xEεx\in\partial E_{\varepsilon} essentially depends on contributions from points yEy\in E in at most two directions.

Using these same ingredients, we establish our second main result regarding the cardinality of the different types of singularities.

Theorem 2.

For any compact set E2E\subset\mathbb{R}^{2}, the number of wedges (S1), sharp singularities (S2, S3 and S8), one-sided shallow singularities (S4) and chain singularities (S6) on Eε\partial E_{\varepsilon} is at most countably infinite.

In addition, we present examples which illustrate that the sets of shallow-shallow singularities (S5) and chain-chain singularities (S7) may be uncountable on Eε\partial E_{\varepsilon}, see Examples 5.5 and 5.7. We refer to the union of categories S6, S7 and S8 as chain singularities and denote the set of chain singularities on Eε\partial E_{\varepsilon} by 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}). Even though the set of chain-chain singularities (S7) may in general be uncountable and can have a positive Hausdorff measure on the boundary Eε\partial E_{\varepsilon}, our third main result establishes the fact that 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) is closed and totally disconnected.

Theorem 3.

For any compact set E2E\subset\mathbb{R}^{2} and ε>0\varepsilon>0, the set 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) of chain singularities is closed and totally disconnected.

As a corollary, Theorem 3 implies that 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) is nowhere dense on Eε\partial E_{\varepsilon}, and is hence small in the topological sense.

1.3. Context

Building on the topological groundwork of [4] and [7], our paper constitutes a first step in the analysis of the local geometry and topological properties of the boundary Eε\partial E_{\varepsilon}, with a particular focus on singularities. A main difference with the majority of the existing literature on boundaries of ε\varepsilon-neighourhoods in this direction [4, 7, 14, 15, 28, 29, 33], is that we do not require (implicit) conditions on ε\varepsilon.

-neighbourhoods arise in many branches of mathematics, ranging from convex analysis and manifold theory [24] to fractal geometry [18] and stochastic processes [20]. In the latter, so-called Wiener sausages represent smoothed-out counterparts of Brownian motion trajectories [10, 16, 25, 26], with applications in theoretical physics [17, 22, 32]. The interplay between the surface area and volume of ε\varepsilon-neighbourhoods [27, 33] and different notions of dimension [18], as well as the dependence of the manifold structure of the boundary Eε\partial E_{\varepsilon} on the radius ε\varepsilon [13, 14, 15, 28, 29] have all received considerable attention during the last decades. In particular, Rataj and collaborators [25, 26, 27, 28, 29] have advanced the understanding in recent years.

The study of ε\varepsilon-neigbourhoods EεE_{\varepsilon} and their boundaries dates back at least to the 1940s and Paul Erdős's remarks [11] regarding their measurability. Following Brown's (1972) initial topological observations [7], Ferry (1976) showed that ε\varepsilon-boundaries are (d1)(d-1)-manifolds in dimensions d=2,3d=2,3 for almost all ε>0\varepsilon>0, but that this fails for d4d\geq 4 [14]. Setting up the problem in a more extensive theoretical framework, Fu (1985) used the semiconcavity of the distance function xdist(x,E)x\mapsto\operatorname{dist}(x,E) to show that the complement dEε\mathbb{R}^{d}\setminus E_{\varepsilon} is a set of positive reach, as defined by Federer (1959) [13, 15]. This allowed him to show that for an arbitrary compact set EdE\subset\mathbb{R}^{d} in dimension d3d\leq 3 there exists an exceptional set RE+R_{E}\subset\mathbb{R}_{+} of Lebesgue measure 0, with the property that the boundaries Eε\partial E_{\varepsilon} are Lipschitz manifolds whenever εRE\varepsilon\notin R_{E}. Rataj and Zajíček (2020) has improved further on these results by providing optimal conditions on the smallness of the set RER_{E} [29].

1.4. Outlook

Our novel technique also allows for the analysis of global topological and regularity properties of the boundary Eε\partial E_{\varepsilon}. This is the topic of the sequel (part II) to this paper. While our results for the moment concern the properties boundaries of ε\varepsilon-neighbourhoods of only planar sets E2E\subset\mathbb{R}^{2}, our techniques appear well-suited for obtaining local and global properties of boundaries of ε\varepsilon-neighbourhoods also in higher dimensions. It would be of particular interest to consider the above-mentioned results of Ferry [14] from this complementary point of view.

Finally, the current paper and its sequel have arisen from our interest in bifurcations of minimal invariant sets of random dynamical systems with bounded noise, which naturally appear as dynamically defined ε\varepsilon-neighbourhoods. In this context, the aim is to develop a theory which allows for the characterisations of topological and/or geometric changes of such sets in parametrised families. The results in this paper provide a characterisation of boundaries at fixed values of parameters (including ε\varepsilon), which is a first step towards more general results concerning the classifications of qualitative changes of minimal invariant sets in (generic) parametrised families of random dynamical systems with bounded noise.

1.5. Structure of the paper

The rest of this paper is structured as follows. In Section 2 we lay out the basic conceptual framework and terminology that will be used throughout the paper. Section 2.2 contains a concise introduction to the notions and basic properties of contributors (Definition 2.1) and outward directions (Definition 2.4) and their relationship with tangential properties of the boundary Eε\partial E_{\varepsilon}.

In Section 3 we shed light on local properties of boundary points xEεx\in\partial E_{\varepsilon} in small neighbourhoods Br(x)¯\overline{B_{r}(x)}. The key result is Proposition 3.1, which states that for E2E\subset\mathbb{R}^{2} the boundary geometry near each xEεx\in\partial E_{\varepsilon} is defined solely by those yEy\in\partial E that lie near the extremal contributors (Definition 2.6) of xx. Building on this insight and a related approximation scheme (Definition 3.3) we show that the boundary can be represented locally by a finite union of continuous graphs (Proposition 3.5). This representation plays a pivotal role in the proofs of subsequent. Section 3.3 contains an analysis of the topological and geometric structure of the complement EεcE_{\varepsilon}^{c} near smooth points, wedges (S1) and shallow singularities (S4–S5).

Sections 4 and 5 contain the main results of this paper. We lay out the different types of singularities encountered on the boundary Eε\partial E_{\varepsilon} (Definition 4.1) and show how various types of singularities can be characterised in terms of the local topological structure of the complement EεcE_{\varepsilon}^{c} and the geometric properties of the boundary Eε\partial E_{\varepsilon} (Propositions 4.2 and 4.5, Corollary 4.3). Section 4 culminates with the proof of our first main result (Theorem 1), by establishing the fact that the classification of singularities, provided in Definition 4.1, defines a partition of Eε\partial E_{\varepsilon}.

Finally, in Section 5.1 cardinalities of the sets of different types of singularities are discussed, and the paper concludes with the proofs of the other two main results, Theorems 2 and  3.

2. -neighbourhoods

The object of our study is the ε\varepsilon-neighbourhood Eε=Bε(E)¯E_{\varepsilon}=\overline{B_{\varepsilon}(E)} of a closed subset EdE\subset\mathbb{R}^{d}. The main results of this paper concern ε\varepsilon-neighbourhoods of planar sets E2E\subset\mathbb{R}^{2}, but we provide the basic definitions in a more general dd-dimensional setting. Throughout the paper we make the assumption that the underlying set EdE\subset\mathbb{R}^{d} is closed111Note that this is not an actual restriction, since dist(x,E)=dist(x,E¯)\mathrm{dist}(x,E)=\mathrm{dist}(x,\overline{E}) for all xdx\in\mathbb{R}^{d} and EdE\subset\mathbb{R}^{d}. and ε>0\varepsilon>0. Many of the results require the stronger assumption of compactness; where necessary, this will be explicitly stated in the formulation of each result.

The most immediate observation regarding the structure of the set EεE_{\varepsilon} is that each xEεx\in\partial E_{\varepsilon} necessarily lies on the boundary of a closed ball Bε(y)¯\overline{B_{\varepsilon}(y)} of radius ε\varepsilon, centered at some yEy\in\partial E. On the other hand, for each xEεx\in\partial E_{\varepsilon} there may exist more than one yEy\in\partial E with yx=ε\left\lVert y-x\right\rVert=\varepsilon. These considerations motivate the following definition.

Definition 2.1 (Contributor).

Let EdE\subset\mathbb{R}^{d} be closed. For each xEεx\in\partial E_{\varepsilon} we define the set of contributors as the collection

ΠE(x):={yE:yx=ε}.\Pi_{E}(x):=\big{\{}y\in\partial E\,\,:\,\,\left\lVert y-x\right\rVert=\varepsilon\big{\}}.

Boundary points xEεx\in\partial E_{\varepsilon} with only one contributor constitute the set

Unpε(E):={xEε:ΠE(x)={y}for someyE},\mathrm{Unp_{\varepsilon}}(E):=\big{\{}x\in\partial E_{\varepsilon}\,:\,\Pi_{E}(x)=\{y\}\,\,\textrm{for some}\,\,y\in\partial E\big{\}},

where Unp\mathrm{Unp} stands for 'unique nearest point', see [13, Definition 4.1].

The set of contributors ΠE(x)\Pi_{E}(x) consists of those points on E\partial E that minimise the distance from E\partial E to xx. Hence ΠE\Pi_{E} can be interpreted as a restriction onto Eε\partial E_{\varepsilon} of the (set-valued) projection projE:2E\mathrm{proj}_{E}:\mathbb{R}^{2}\to E given by projE(x):={yE:yx=dist(x,E)}\mathrm{proj}_{E}(x):=\{y\in E\,:\,\left\lVert y-x\right\rVert=\mathrm{dist}(x,E)\}. In terms of our classification of boundary points (see Definition 4.1 and Figures 1 and 9) the set Unpε(E)\mathrm{Unp_{\varepsilon}}(E) consists of smooth points and shallow singularities (S4–S5).

Definition 2.2 (Smooth point, singularity).

We call a boundary point xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E) smooth, if there exists a neighbourhood Br(x)B_{r}(x) for which EεBr(x)Unpε(E)\partial E_{\varepsilon}\cap B_{r}(x)\subset\mathrm{Unp_{\varepsilon}}(E). If xx is not smooth, we call it a singularity and write xS(Eε)x\in S(E_{\varepsilon}).

The rationale for Definition 2.2 stems from the fact that any smooth xEεx\in\partial E_{\varepsilon} in terms of Definition 2.2 turns out to be equivalent to xx having a neighbourhood Br(x)B_{r}(x) in which the boundary is a C1C^{1}-smooth curve, see Proposition 4.6.

2.1. Tangents via Outward Directions

Our first objective is to shed light on the tangential properties of individual boundary points xEεx\in\partial E_{\varepsilon}. Acknowledging that classical tangents do not necessarily exist everywhere on the boundary, we adopt a set-valued definition of tangency which allows for several tangential directions to exist at each point. Our definition is a restriction of [13, Definition 4.3] to the boundary Eε\partial E_{\varepsilon}.

Definition 2.3 (Tangent set).

Let EdE\subset\mathbb{R}^{d} be closed and xEεx\in\partial E_{\varepsilon}. We define the set Tx(Eε)T_{x}(E_{\varepsilon}) of unit tangent vectors of EεE_{\varepsilon} at xx as all those points vSd1v\in S^{d-1} for which there exists a sequence (xn)n=1E(x_{n})_{n=1}^{\infty}\subset\partial E of boundary points satisfying xnxx_{n}\to x and

xnxxnxv,as n.\frac{x_{n}-x}{\left\lVert x_{n}-x\right\rVert}\rightarrow v,\quad\textrm{as }\,n\to\infty.

In order to study the existence of tangential directions at boundary points xEεx\in\partial E_{\varepsilon}, we relate the set Tx(Eε)T_{x}(E_{\varepsilon}) to what we call outward directions. Intuitively, the set of outward directions at each xx contains the angles at which xx can be approached from the complement Eεc:=dEεE_{\varepsilon}^{c}:=\mathbb{R}^{d}\setminus E_{\varepsilon}. It turns out that for an ε\varepsilon-neighbourhood EεE_{\varepsilon}, the extremal values of these angles coincide with the tangential directions as defined in Definition 2.3. Hence the existence of tangents at each xEεx\in\partial E_{\varepsilon} hinges on the existence and properties of corresponding outward directions, which turn out to be easier to study due to their geometric relationship with the contributors yΠE(x)y\in\Pi_{E}(x).

We define outward directions as points on the unit sphere Sd1dS^{d-1}\subset\mathbb{R}^{d} but think of them rather as directional vectors in the ambient space d\mathbb{R}^{d}, since we want to operate with them using the Euclidean scalar product ,:d×d\langle\cdot,\cdot\rangle\,:\,\mathbb{R}^{d}\times\mathbb{R}^{d}\to\mathbb{R}.

Definition 2.4 (Outward direction).

Let EdE\subset\mathbb{R}^{d} be closed. We say that a point ξSd1\xi\in S^{d-1} is an outward direction from EεE_{\varepsilon} at a boundary point xEεx\in\partial E_{\varepsilon}, if there exists a sequence (xn)n=1Eεc(x_{n})_{n=1}^{\infty}\subset E_{\varepsilon}^{c}, for which xnxx_{n}\to x and

ξn:=xnxxnxξSd1d,\xi_{n}:=\frac{x_{n}-x}{\left\lVert x_{n}-x\right\rVert}\longrightarrow\xi\in S^{d-1}\subset\mathbb{R}^{d},

as nn\to\infty. We denote by Ξx(Eε)\Xi_{x}(E_{\varepsilon}) the set of outward directions from EεE_{\varepsilon} at xx.

Refer to caption
(a) A singularity xEεx\in\partial E_{\varepsilon} with ΠEext(x)={y1,y2}\Pi_{E}^{\mathrm{ext}}(x)=\{y_{1},y_{2}\} and Ξxext(Eε)={ξ1,ξ2}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi_{1},\xi_{2}\}.
Refer to caption
(b) The set Ξx(Eε)S1\Xi_{x}(E_{\varepsilon})\subset S^{1} of outward directions is geodesically convex with boundary S1Ξx(Eε)={ξ1,ξ2}\partial_{S^{1}}\Xi_{x}(E_{\varepsilon})=\{\xi_{1},\xi_{2}\}.
Figure 2. Illustration of the relationship between outward directions and extremal contributors.
Remark 2.5.

The concept of outward directions is a variation of the well-known contingent cone, introduced by Bouligand (see for instance [2, 3, 31] and Bouligand's original work [5, 6]). For a boundary point xEεx\in\partial E_{\varepsilon} the contingent cone (Bouligand cone) Cx(Eε)C_{x}(E_{\varepsilon}) consists of those vectors vdv\in\mathbb{R}^{d}, for which there exist sequences (hn)n=1+(h_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} and (vn)n=1d(v_{n})_{n=1}^{\infty}\subset\mathbb{R}^{d} for which x+hnvnEεx+h_{n}v_{n}\in E_{\varepsilon} for all nn\in\mathbb{N} and

hn0,andvnvh_{n}\to 0,\quad\textrm{and}\quad v_{n}\to v

as nn\to\infty. If instead of the outward directions Ξx(Eε)\Xi_{x}(E_{\varepsilon}) one considers at each xEεx\in\partial E_{\varepsilon} the contingent cone Cx(Eεc)C_{x}(E_{\varepsilon}^{c}) for the complement EεcE_{\varepsilon}^{c}, it follows that 𝒲x(Eε)=Cx(Eεc)\mathcal{W}_{x}(E_{\varepsilon})=C_{x}(E_{\varepsilon}^{c}), where

𝒲x(Eε):={sξ:ξΞx(Eε),s0}\mathcal{W}_{x}(E_{\varepsilon}):=\left\{s\xi\,:\,\xi\in\Xi_{x}(E_{\varepsilon}),\,s\geq 0\right\}

denotes the outward cone at xx. We do not make use of this correspondence, but for further information on tangent cones, see for instance [8, 30].

For each xEεx\in\partial E_{\varepsilon} we single out those outward directions ξΞx(Eε)\xi\in\Xi_{x}(E_{\varepsilon}) that are perpendicular to some contributor yΠE(x)y\in\Pi_{E}(x)—we call these the extremal outward directions and extremal contributors (see Figure 2). Definition 2.6 below emphasises this geometric relationship, while Proposition 2.12 in the next subsection confirms that extremal outward directions can equivalently be defined via the topological property of constituting the boundary of the set of outward directions Ξx(Eε)\Xi_{x}(E_{\varepsilon}).

Definition 2.6 (Extremal contributor, extremal outward direction).

Let EdE\subset\mathbb{R}^{d} be closed and xEεx\in\partial E_{\varepsilon}. If an outward direction ξΞx(Eε)\xi\in\Xi_{x}(E_{\varepsilon}) and a contributor yΠE(x)y\in\Pi_{E}(x) satisfy

yx,ξ=0,\langle y-x,\xi\rangle=0,

we call ξ\xi an extremal outward direction and yy an extremal contributor at xx. For each xEεx\in\partial E_{\varepsilon}, we write Ξxext(Eε)\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) and ΠEext(x)\Pi_{E}^{\mathrm{ext}}(x) for the sets of extremal outward directions and extremal contributors, respectively.

The precise correspondence between extremal contributors, extremal outward directions, and tangential directions at each xEεx\in\partial E_{\varepsilon} is presented in Section 2.2, where we collect in one place all the basic results that we need in the remainder of the paper. The existence of outward directions at each xEεx\in\partial E_{\varepsilon} is established in Proposition 2.7 and their geometric relationship with the contributors is explored in Lemma 2.10 and Proposition 2.12. The coincidence of the set of tangential directions Tx(E)T_{x}(E) with the set of extremal outward directions Ξxext(Eε)\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) is established in Proposition 2.14.

2.2. Properties of Contributors and Outward Directions

We collect here the basic properties of contributors and outward directions that we need in our analysis of the boundary Eε\partial E_{\varepsilon}. The proofs make repeated use of convergent subsequences and scalar products and are rather elementary, although at places somewhat tedious. As before, the set EdE\subset\mathbb{R}^{d} is assumed to be closed, and ε>0\varepsilon>0.

Proposition 2.7 (The set of outward directions is non-empty and closed).

Let EdE\subset\mathbb{R}^{d} be closed and xEεx\in\partial E_{\varepsilon}. Then the set Ξx(Eε)\Xi_{x}(E_{\varepsilon}) of outward directions is non-empty and closed.

Proof.

Let xEεx\in\partial E_{\varepsilon} and choose some sequence (xn)n=1(x_{n})_{n=1}^{\infty} in EεcE_{\varepsilon}^{c} with xnxx_{n}\to x. This implies xnxx_{n}\neq x for all nn\in\mathbb{N} since EεE_{\varepsilon} is closed. One may hence define a sequence (ξn)n=1(\xi_{n})_{n=1}^{\infty} in S1S^{1} by setting ξn:=(xnx)/xnx\xi_{n}:=(x_{n}-x)/\left\lVert x_{n}-x\right\rVert for all nn\in\mathbb{N}. The compactness of S1S^{1} implies that (ξn)n=1(\xi_{n})_{n=1}^{\infty} has a convergent subsequence, the limit of which is an element in Ξx(Eε)\Xi_{x}(E_{\varepsilon}).

To show that Ξx(Eε)\Xi_{x}(E_{\varepsilon}) is closed, let ξSd1\xi\in S^{d-1} and assume there exists a sequence (ξ(n))n=1Ξx(Eε)\left(\xi^{(n)}\right)_{n=1}^{\infty}\subset\Xi_{x}(E_{\varepsilon}) with ξ(n)ξ\xi^{(n)}\to\xi as nn\to\infty. One needs to show that this implies ξΞx(Eε)\xi\in\Xi_{x}(E_{\varepsilon}). We first use Definition 2.4 to identify each of the directions ξ(n)\xi^{(n)} with a convergent sequence in EεcE_{\varepsilon}^{c}, and then apply a kind of diagonalisation argument in order to construct a new sequence (zn)n=1(z_{n})_{n=1}^{\infty} in EεcE_{\varepsilon}^{c} with (znx)/znxξ(z_{n}-x)/\left\lVert z_{n}-x\right\rVert\to\xi.

Without loss of generality, let ξ(n)ξ1n\left\lVert\xi^{(n)}-\xi\right\rVert\leq\frac{1}{n} for all nn\in\mathbb{N}. Now, for each nn\in\mathbb{N} one can choose a sequence (xk(n))k=1Eεc\big{(}x_{k}^{(n)}\big{)}_{k=1}^{\infty}\subset E_{\varepsilon}^{c} with xk(n)xx_{k}^{(n)}\to x and

ξk(n):=xk(n)xxk(n)xξ(n),as k.\xi_{k}^{(n)}:=\frac{x_{k}^{(n)}-x}{\big{\|}x_{k}^{(n)}-x\big{\|}}\longrightarrow\xi^{(n)},\quad\textrm{as }\,k\to\infty.

Consequently there exists for each nn\in\mathbb{N} some K(n)K(n)\in\mathbb{N}, for which xk(n)x1/n\big{\|}x_{k}^{(n)}-x\big{\|}\leq 1/n and ξk(n)ξ(n)1/n\big{\|}\xi_{k}^{(n)}-\xi^{(n)}\big{\|}\leq 1/n for all kK(n)k\geq K(n). Using these indices one can define a new sequence by setting zn:=xK(n)(n)z_{n}:=x_{K(n)}^{(n)} for each nn\in\mathbb{N}. Accordingly

znx=xK(n)(n)x1n\left\lVert z_{n}-x\right\rVert=\big{\|}x_{K(n)}^{(n)}-x\big{\|}\leq\frac{1}{n}

so that znxz_{n}\to x. Furthermore, writing ξnz:=(znx)/znx=ξK(n)(n)\xi_{n}^{z}:=(z_{n}-x)/\left\lVert z_{n}-x\right\rVert=\xi_{K(n)}^{(n)}, we have

ξnzξ\displaystyle\left\lVert\xi_{n}^{z}-\xi\right\rVert ξK(n)(n)ξ(n)+ξ(n)ξ2n0\displaystyle\leq\big{\|}\xi_{K(n)}^{(n)}-\xi^{(n)}\big{\|}+\big{\|}\xi^{(n)}-\xi\big{\|}\leq\frac{2}{n}\longrightarrow 0

as nn\to\infty. Thus ξ\xi is the outward direction corresponding to the sequence (zn)n=1\left(z_{n}\right)_{n=1}^{\infty}, which implies ξΞx(Eε)\xi\in\Xi_{x}(E_{\varepsilon}), as required. ∎

Despite their simplicity, the following Lemmas 2.8 and 2.9 regarding contributors are a key ingredient in many of the subsequent proofs.

Lemma 2.8 (Convergence of contributors).

Let EdE\in\mathbb{R}^{d} be compact, let (xn)n=1Eε(x_{n})_{n=1}^{\infty}\subset E_{\varepsilon} with xnxEεx_{n}\to x\in\partial E_{\varepsilon}, and (yn)n=1E(y_{n})_{n=1}^{\infty}\subset E with xnBε(yn)¯x_{n}\in\overline{B_{\varepsilon}(y_{n})} for all nn\in\mathbb{N}. Then there exists some yΠE(x)y\in\Pi_{E}(x) and a convergent subsequence (ynk)k=1(y_{n_{k}})_{k=1}^{\infty}, for which ynkyy_{n_{k}}\to y as kk\to\infty.

Proof.

Due to compactness of EE, there exists a convergent subsequence (ynk)k=1(y_{n_{k}})_{k=1}^{\infty} with ynkyEy_{n_{k}}\to y\in E as kk\to\infty. For each kk\in\mathbb{N},

yxyynk+ynkxnk+xnkx,\left\lVert y-x\right\rVert\leq\left\lVert y-y_{n_{k}}\right\rVert+\left\lVert y_{n_{k}}-x_{n_{k}}\right\rVert+\left\lVert x_{n_{k}}-x\right\rVert,

which implies yxε\left\lVert y-x\right\rVert\leq\varepsilon, since xnkxx_{n_{k}}\to x, ynkyy_{n_{k}}\to y, and limkynkxnkε\lim_{k\to\infty}\left\lVert y_{n_{k}}-x_{n_{k}}\right\rVert\leq\varepsilon. On the other hand yxdist(x,E)=ε\left\lVert y-x\right\rVert\geq\textrm{dist}(x,E)=\varepsilon. Hence yx=ε\left\lVert y-x\right\rVert=\varepsilon, so that yΠE(x)y\in\Pi_{E}(x). ∎

Lemma 2.9 (Tails of directed sequences).

Let EdE\subset\mathbb{R}^{d} be compact, (xn)n=1d(x_{n})_{n=1}^{\infty}\subset\mathbb{R}^{d} with xnxEεx_{n}\to x\in\partial E_{\varepsilon} and xnxx_{n}\neq x for all nn\in\mathbb{N}, and assume

vn:=xnxxnxv,as n.v_{n}:=\frac{x_{n}-x}{\left\lVert x_{n}-x\right\rVert}\longrightarrow v,\quad\textrm{as }\,n\to\infty.
  1. (i)

    If yx,v>0\langle y-x,v\rangle>0 for some yΠE(x)y\in\Pi_{E}(x), then there exists some NN\in\mathbb{N}, for which xnBε(y)x_{n}\in B_{\varepsilon}(y) for all nNn\geq N.

  2. (ii)

    If yx,v<0\langle y-x,v\rangle<0 for all yΠE(x)y\in\Pi_{E}(x), then there exists some NN\in\mathbb{N}, for which xnEεcx_{n}\in E_{\varepsilon}^{c} for all nNn\geq N.

Proof.

(i) Assume yx,v=p>0\langle y-x,v\rangle=p>0 for some yΠE(x)y\in\Pi_{E}(x). Then yx,vnp\langle y-x,v_{n}\rangle\to p due to the continuity of the scalar product. Hence there exists some NN\in\mathbb{N} for which yx,vn>12p\langle y-x,v_{n}\rangle>\frac{1}{2}p and xnx<p\left\lVert x_{n}-x\right\rVert<p, whenever nNn\geq N. This implies

xnx<p<2yx,vn\left\lVert x_{n}-x\right\rVert<p<2\langle y-x,v_{n}\rangle

for all nNn\geq N so that

yxn2\displaystyle\left\lVert y-x_{n}\right\rVert^{2} =(yx)(xnx)2\displaystyle=\left\lVert(y-x)-(x_{n}-x)\right\rVert^{2}
=ε2+xnx(xnx2yx,vn)<ε2.\displaystyle=\varepsilon^{2}+\left\lVert x_{n}-x\right\rVert\left(\left\lVert x_{n}-x\right\rVert-2\langle y-x,v_{n}\rangle\right)<\varepsilon^{2}.

(ii) Assume to the contrary that there exists a subsequence (xnk)k=1Eε(x_{n_{k}})_{k=1}^{\infty}\subset E_{\varepsilon}. Then for each kk there exists some ykEy_{k}\in E for which xnkykε\left\lVert x_{n_{k}}-y_{k}\right\rVert\leq\varepsilon. Since EE is compact, Lemma 2.8 implies the existence of some yΠE(x)y^{*}\in\Pi_{E}(x) for which ykyy_{k}\to y^{*} (if necessary, one can switch to a further convergent subsequence). By assumption yΠE(x)y^{*}\in\Pi_{E}(x) implies yx,v<0\langle y^{*}-x,v\rangle<0. Due to the continuity of the scalar product there exists some q>0q>0 and KK\in\mathbb{N}, for which ykxnk,vnk12q\langle y_{k}-x_{n_{k}},v_{n_{k}}\rangle\leq-\frac{1}{2}q and xnkx<q\left\lVert x_{n_{k}}-x\right\rVert<q, whenever kKk\geq K. Then

xnkx+2ykxnk,vnk<0\left\lVert x_{n_{k}}-x\right\rVert+2\langle y_{k}-x_{n_{k}},v_{n_{k}}\rangle<0

for all kKk\geq K, and applying the triangle-inequality with respect to the points xnkx_{n_{k}} yields

ykx2\displaystyle\left\lVert y_{k}-x\right\rVert^{2} ε2+xnkx(xnkx+2ykxnk,vnk)<ε2.\displaystyle\leq\varepsilon^{2}+\left\lVert x_{n_{k}}-x\right\rVert\left(\left\lVert x_{n_{k}}-x\right\rVert+2\langle y_{k}-x_{n_{k}},v_{n_{k}}\rangle\right)<\varepsilon^{2}.

This implies the contradiction xint(Eε)x\in\textrm{int}(E_{\varepsilon}). ∎

Lemma 2.10 below provides a partial characterisation of outward directions Ξx(Eε)\Xi_{x}(E_{\varepsilon}) in terms of the contributors ΠE(x)\Pi_{E}(x). Geometrically it implies that outward directions point away from the vectors yxy-x for all yΠE(x)y\in\Pi_{E}(x), see Figure 2.

Lemma 2.10 (Orientation of outward directions relative to contributors).

Let EdE\subset\mathbb{R}^{d} be compact, xEεx\in\partial E_{\varepsilon} and ξSd1\xi\in S^{d-1}. Then

  1. (i)

    if ξΞx(Eε)\xi\in\Xi_{x}(E_{\varepsilon}), then yx,ξ0\langle y-x,\xi\rangle\leq 0 for all yΠE(x)y\in\Pi_{E}(x),

  2. (ii)

    if yx,ξ<0\langle y-x,\xi\rangle<0 for all yΠE(x)y\in\Pi_{E}(x), then ξΞx(Eε)\xi\in\Xi_{x}(E_{\varepsilon}).

Proof.

(i) If ξΞx(Eε)\xi\in\Xi_{x}(E_{\varepsilon}), there exists a sequence (xn)n=1Eεc(x_{n})_{n=1}^{\infty}\subset E_{\varepsilon}^{c}, for which xnxx_{n}\to x and ξn:=(xnx)/xnxξ\xi_{n}:=(x_{n}-x)/\left\lVert x_{n}-x\right\rVert\to\xi as nn\to\infty. Assume contrary to the claim tha there exists some yΠE(x)y\in\Pi_{E}(x) with yx,ξ>0\langle y-x,\xi\rangle>0. Substituting vn=ξnv_{n}=\xi_{n} and v=ξv=\xi in Lemma 2.9 (i) implies the existence of some NN\in\mathbb{N} for which xnint(Eε)x_{n}\in\textrm{int}(E_{\varepsilon}) for all nNn\geq N. This contradicts the claim.

(ii) Write ξn=1nξ\xi_{n}=\frac{1}{n}\xi, and define xn:=x+ξnx_{n}:=x+\xi_{n}, so that (xnx)/xnx=ξ(x_{n}-x)/\left\lVert x_{n}-x\right\rVert=\xi for all nn\in\mathbb{N}. Substituting v:=ξv:=\xi and vn:=ξnv_{n}:=\xi_{n} in Lemma 2.9 (ii) implies the existence of some NN\in\mathbb{N} for which xnEεcx_{n}\in E_{\varepsilon}^{c} for all nNn\geq N. The sequence (xn)n=N(x_{n})_{n=N}^{\infty} now defines the outward direction ξΞx(Eε)\xi\in\Xi_{x}(E_{\varepsilon}). ∎

Note that assuming the weaker condition yx,ξ0\langle y-x,\xi\rangle\leq 0 for all contributors yΠE(x)y\in\Pi_{E}(x) is not sufficient in Lemma 2.10 (ii). For example, let E:=[2,3]×{0,1}E:=[2,3]\times\{0,1\} and consider the set EεE_{\varepsilon} with ε=1/2\varepsilon=1/2. Then x:=(3,1/2)Eεx:=(3,1/2)\in\partial E_{\varepsilon} with Π(x)={(3,0),(3,1)}\Pi(x)=\{(3,0),(3,1)\} and has only one outward direction ξ=(1,0)\xi=(1,0). Here also η:=(1,0)\eta:=(-1,0) satisfies yx,η=0\langle y-x,\eta\rangle=0 for y{(3,0),(3,1)}y\in\{(3,0),(3,1)\}, and yet ηΞx(Eε)\eta\notin\Xi_{x}(E_{\varepsilon}). This example illustrates the difference between Eε\partial E_{\varepsilon} and the ε\varepsilon-boundary E<ε\partial E_{<\varepsilon} (see Section 1.1), since here E<εEε=(2,3)×{1/2}intEε\partial E_{<\varepsilon}\setminus\partial E_{\varepsilon}=(2,3)\times\{1/2\}\subset\mathrm{int}\,E_{\varepsilon}. See also [27, Example 2.1].

In order to describe the geometry of the sets of outward directions Ξx(Eε)\Xi_{x}(E_{\varepsilon}) on the circle S1S^{1}, we introduce the concept of a geodesic arc-segment. Intuitively, a geodesic arc-segment is the shortest curve on S1S^{1} that connects two points v,wS1v,w\in S^{1}.

Definition 2.11 (Geodesic arc-segment).

Let v,wS12v,w\in S^{1}\subset\mathbb{R}^{2} and let

(2.1) [v,w]S1:={uS1:u=av+bw for some a,b0}.[v,w]_{S^{1}}:=\left\{u\in S^{1}\,:\,u=av+bw\,\textrm{ for some }a,b\geq 0\right\}.

For wvw\neq-v, the set [v,w]S1[v,w]_{S^{1}} defines a geodesic arc-segment between vv and ww. We also define the corresponding open geodesic arc-segment (v,w)S1S1(v,w)_{S^{1}}\subset S^{1} as

(2.2) (v,w)S1:=[v,w]S1{v,w}.(v,w)_{S^{1}}:=[v,w]_{S^{1}}\setminus\{v,w\}.

We use the notations [v,w]S1[v,w]_{S^{1}} and (v,w)S1(v,w)_{S^{1}} in accordance with (2.1) and (2.2) also for the cases v=wv=w and v=wv=-w, even though the corresponding sets in these cases are not arc-segments.

Unlike the previous results in this section, we formulate and prove the statements in Proposition 2.12 and Lemma 2.13 below only for the two-dimensional case. Note also that Proposition 2.12 is formulated for a compact set E2E\subset\mathbb{R}^{2}, but essentially the same proof works for any closed set EE due to the local nature of the result.

Proposition 2.12 (Structure of sets of outward directions).

Let E2E\subset\mathbb{R}^{2} be compact and xEεx\in\partial E_{\varepsilon}. Then the set of outward directions Ξx(Eε)\Xi_{x}(E_{\varepsilon}) satisfies the following.

  1. (i)

    If xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E), then Ξx(Eε)={ξS1:yx,ξ0}\Xi_{x}(E_{\varepsilon})=\left\{\xi\in S^{1}\,:\,\langle y-x,\xi\rangle\leq 0\right\};

  2. (ii)

    If xUnpε(E)x\notin\mathrm{Unp_{\varepsilon}}(E), then Ξx(Eε)=[ξ1,ξ2]S1\Xi_{x}(E_{\varepsilon})=[\xi_{1},\xi_{2}]_{S^{1}}, where ξ1,ξ2\xi_{1},\xi_{2} are the only extremal outward directions at xx, possibly satisfying ξ1=ξ2\xi_{1}=\xi_{2}.

Proof.

(i) Since ΠE(x)={y}\Pi_{E}(x)=\{y\}, Lemma 2.10 (ii) implies

X:={ξS1:yx,ξ<0}Ξx(Eε).X:=\{\xi\in S^{1}\,:\,\langle y-x,\xi\rangle<0\}\subset\Xi_{x}(E_{\varepsilon}).

Then S1X={ξS1:yx,ξ=0}\partial_{S^{1}}X=\{\xi\in S^{1}\,:\,\langle y-x,\xi\rangle=0\} due to continuity of the scalar product, and Lemmas 2.7 and 2.10 (i) imply X¯=Ξx(Eε)\overline{X}=\Xi_{x}(E_{\varepsilon}), as claimed.

(ii) Assume then that ΠE(x)\Pi_{E}(x) contains at least two points. We assert that

  1. (a)

    if ξ,ηΞx(Eε)\xi,\eta\subset\Xi_{x}(E_{\varepsilon}) and γ(ξ,η)S1\gamma\in(\xi,\eta)_{S^{1}}, then yx,γ<0\langle y-x,\gamma\rangle<0 for all yΠE(x)y\in\Pi_{E}(x),

  2. (b)

    ξintS1Ξx(Eε)\xi\in\textrm{int}_{S^{1}}\Xi_{x}(E_{\varepsilon}) if and only if yx,ξ<0\langle y-x,\xi\rangle<0 for all yΠE(x)y\in\Pi_{E}(x),

  3. (c)

    Ξx(Eε)=[ξ1,ξ2]S1\Xi_{x}(E_{\varepsilon})=[\xi_{1},\xi_{2}]_{S^{1}}.

(a) If Ξx(Eε)={ξ}\Xi_{x}(E_{\varepsilon})=\{\xi\} or Ξx(Eε)={ξ,ξ}\Xi_{x}(E_{\varepsilon})=\{\xi,-\xi\} for some ξS1\xi\in S^{1}, the claim is true since (ξ,ξ)S1=(ξ,ξ)S1=(\xi,\xi)_{S^{1}}=(\xi,-\xi)_{S^{1}}=\varnothing.

Let ξ,ηΞx(Eε)\xi,\eta\in\Xi_{x}(E_{\varepsilon}) with η{ξ,ξ}\eta\notin\{\xi,-\xi\} and define a parametrised curve γ:[0,1]S1\gamma:[0,1]\to S^{1} by

γ(t):=tη+(1t)ξtη+(1t)ξ.\gamma(t):=\frac{t\eta+(1-t)\xi}{\left\lVert t\eta+(1-t)\xi\right\rVert}.

Clearly γ(0)=ξ,γ(1)=η\gamma(0)=\xi,\gamma(1)=\eta, and γ((0,1))=(ξ,η)S1\gamma((0,1))=(\xi,\eta)_{S^{1}}. For each yΠE(x)y\in\Pi_{E}(x), consider the scalar product

(2.3) Py(t):=yx,tη+(1t)ξ\displaystyle P_{y}(t):=\langle y-x,t\eta+(1-t)\xi\rangle =tyx,η+(1t)yx,ξ\displaystyle=t\langle y-x,\eta\rangle+(1-t)\langle y-x,\xi\rangle
(2.4) =yx,ξ+tyx,ηξ.\displaystyle=\langle y-x,\xi\rangle+t\langle y-x,\eta-\xi\rangle.

We show that Py(t)<0P_{y}(t)<0 for every yΠE(x)y\in\Pi_{E}(x) and all t(0,1)t\in(0,1). We have Py(t)0P_{y}(t)\leq 0 for all t[0,1]t\in[0,1] and all yΠE(x)y\in\Pi_{E}(x), since Lemma 2.10 (i) guarantees

yx,η0andyx,ξ0\langle y-x,\eta\rangle\leq 0\quad\textrm{and}\quad\langle y-x,\xi\rangle\leq 0

for all yΠE(x)y\in\Pi_{E}(x). For t(0,1)t\in(0,1), equation (2.3) implies Py(t)<0P_{y}(t)<0 when yx,ξ<0\langle y-x,\xi\rangle<0. On the other hand, if yx,ξ=0\langle y-x,\xi\rangle=0, equation (2.4) implies

tyx,ηξ=Py(t)0.t\langle y-x,\eta-\xi\rangle=P_{y}(t)\leq 0.

Hence the inequality Py(t)<0P_{y}(t)<0 holds if and only if η{ξ,ξ}\eta\notin\{\xi,-\xi\}. This shows that Py(t)<0P_{y}(t)<0 for arbitrary yΠE(x)y\in\Pi_{E}(x), whenever t(0,1)t\in(0,1).

(b) If ξS1\xi\in S^{1} satisfies yx,ξ<0\langle y-x,\xi\rangle<0 for all yΠE(x)y\in\Pi_{E}(x), then there exists some nn\in\mathbb{N} for which yx,ξ<1/n\langle y-x,\xi\rangle<-1/n for all yΠE(x)y\in\Pi_{E}(x). To show this, assume to the contrary that for each nn\in\mathbb{N} there exists some ynΠE(x)y_{n}\in\Pi_{E}(x) for which

ynx,ξ1/n.\langle y_{n}-x,\xi\rangle\geq-1/n.

Since S1S^{1} is compact and EE is closed, there exists a convergent subsequence (ynk)k=1(yn)n=1(y_{n_{k}})_{k=1}^{\infty}\subset(y_{n})_{n=1}^{\infty} with ynkyΠE(x)y_{n_{k}}\to y^{*}\in\Pi_{E}(x). On the other hand, due to the continuity of the scalar product we have

yx,ξ=limkynkx,ξ0,\langle y^{*}-x,\xi\rangle=\lim_{k\to\infty}\langle y_{n_{k}}-x,\xi\rangle\geq 0,

which contradicts the assumption that yx,ξ<0\langle y-x,\xi\rangle<0 for all yΠE(x)y\in\Pi_{E}(x).

Assume now that yx,ξ<0\langle y-x,\xi\rangle<0 for all yΠE(x)y\in\Pi_{E}(x) and that nn\in\mathbb{N} has been chosen so that yx,ξ<1/n\langle y-x,\xi\rangle<-1/n for all yΠE(x)y\in\Pi_{E}(x). The continuity of the scalar product implies that for some δ\delta, depending on nn, one has yx,ξ^<0\left\langle y-x,\widehat{\xi}\right\rangle<0 for all ξ^\widehat{\xi} that satisfy ξ^ξ<δ\left\lVert\widehat{\xi}-\xi\right\rVert<\delta. Hence ξ\xi has an open neighbourhood Bδ(ξ)B_{\delta}(\xi) satisfying Bδ(ξ)S1Ξx(Eε)B_{\delta}(\xi)\cap S^{1}\subset\Xi_{x}(E_{\varepsilon}), which implies ξintS1Ξx(Eε)\xi\in\textrm{int}_{S^{1}}\Xi_{x}(E_{\varepsilon}).

For the other direction, assume ξintS1Ξx(Eε)\xi\in\textrm{int}_{S^{1}}\Xi_{x}(E_{\varepsilon}). Then there exist η1,η2Ξx(Eε)\eta_{1},\eta_{2}\in\Xi_{x}(E_{\varepsilon}), for which ξ(η1,η2)S1intS1Ξx(Eε)\xi\in(\eta_{1},\eta_{2})_{S^{1}}\subset\textrm{int}_{S^{1}}\Xi_{x}(E_{\varepsilon}). Step (a) consequently implies yx,ξ<0\langle y-x,\xi\rangle<0 for all yΠE(x)y\in\Pi_{E}(x).

(c) It follows from steps (a) and (b) that intS1Ξx(Eε)=(ξ1,ξ2)S1\textrm{int}_{S^{1}}\Xi_{x}(E_{\varepsilon})=(\xi_{1},\xi_{2})_{S^{1}}. This in turn implies Ξxext(Eε)={ξ1,ξ2}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi_{1},\xi_{2}\}, when intS1Ξx(Eε)\textrm{int}_{S^{1}}\Xi_{x}(E_{\varepsilon})\neq\varnothing, and Ξx(Eε)={ξ}\Xi_{x}(E_{\varepsilon})=\{\xi\} (singleton), when ξ1=ξ2\xi_{1}=\xi_{2}.∎

Proposition 2.12 thus gives the following geometric picture of the set of extremal outward directions. In the case of a sharp singularity (S2) or a chain singularity (S6) (see Figure 1 and Definition 4.1), the set of extremal outward directions is a singleton Ξxext(Eε)={ξ}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi\} for some ξS1\xi\in S^{1}. Otherwise Ξxext(Eε)\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) contains two points, which may point directly away from each other or form an acute or obtuse angle.

Lemma 2.13 below summarises the limiting behaviour of outward directions ξn\xi_{n} and contributors yny_{n} of points xnx_{n} that appear in convergent sequences on the ε\varepsilon-neighbourhood boundary. In particular, Lemma 2.13 (ii)(a) establishes that for each xEεx\in\partial E_{\varepsilon} the set of tangent vectors Tx(Eε)T_{x}(E_{\varepsilon}) is a subset of the set Ξxext(Eε)\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) of extremal outward directions. According to Proposition 2.14 these sets in fact coincide for all xEεx\in\partial E_{\varepsilon}.

Lemma 2.13 (Orientation in converging sequences of boundary points).

Let E2E\subset\mathbb{R}^{2} be compact and let xEεx\in\partial E_{\varepsilon}. Furthermore, let (xn)n=1(x_{n})_{n=1}^{\infty} be a sequence on Eε\partial E_{\varepsilon} with xnxx_{n}\to x and define ξn:=(xnx)/xnx\xi_{n}:=(x_{n}-x)/\left\lVert x_{n}-x\right\rVert for all nn\in\mathbb{N}. Then the following statements hold true:

  1. (i)

    The sequence (ξn)n=1(\xi_{n})_{n=1}^{\infty} can be split into two disjoint, convergent subsequences (ξi,k)k=1(\xi_{i,k})_{k=1}^{\infty}, where i{1,2}i\in\{1,2\} and ξ(i):=limkξi,kΞxext(Eε)\xi^{(i)}:=\lim_{k\to\infty}\xi_{i,k}\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}).

  2. (ii)

    If the limit ξ:=limnξnS1\xi:=\lim_{n\to\infty}\xi_{n}\in S^{1} exists, then

    1. (a)

      every sequence (yn)n=1(y_{n})_{n=1}^{\infty} in EE with ynΠ(xn)y_{n}\in\Pi(x_{n}) for all nn\in\mathbb{N} has a convergent subsequence (ynk)k=1(y_{n_{k}})_{k=1}^{\infty} for which y:=limkynkΠEext(x)y:=\lim_{k\to\infty}y_{n_{k}}\in\Pi_{E}^{\mathrm{ext}}(x). Furthermore yx,ξ=0\langle y-x,\xi\rangle=0 and consequently ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}).

    2. (b)

      every sequence (ηn)n=1(\eta_{n})_{n=1}^{\infty} in S1S^{1} with ηnΞxnext(Eε)\eta_{n}\in\Xi_{x_{n}}^{\mathrm{ext}}(E_{\varepsilon}) for all nn\in\mathbb{N} satisfies

      limnηn,ξ=1.\lim_{n\to\infty}\left\lVert\langle\eta_{n},\xi\rangle\right\rVert=1.
Proof.

(ii)(a) Due to Lemma 2.8, there exists some yΠE(x)y\in\Pi_{E}(x) and a convergent subsequence (ynk)k=1(yn)n=1(y_{n_{k}})_{k=1}^{\infty}\subset(y_{n})_{n=1}^{\infty}, for which ynkyy_{n_{k}}\to y. We break the proof into three steps.

Step 1. yx,ξ0\langle y-x,\xi\rangle\leq 0: Assume contrary to the claim that yx,ξ>0\langle y-x,\xi\rangle>0. Then substituting xk:=xnkx_{k}:=x_{n_{k}} and v:=ξv:=\xi in Lemma 2.9 (1) implies that for some KK\in\mathbb{N} we have xnkint(Eε)x_{n_{k}}\in\textrm{int}(E_{\varepsilon}) for all kKk\geq K. This contradicts the assumption xnEεx_{n}\in\partial E_{\varepsilon} for all nn\in\mathbb{N}.

Step 2. yx,ξ0\langle y-x,\xi\rangle\geq 0: Assume contrary to the claim that yx,ξ<0\langle y-x,\xi\rangle<0. The continuity of the scalar product then implies that there exists some KK\in\mathbb{N} for which xnkx+2ynkxnk,ξnk<0\left\lVert x_{n_{k}}-x\right\rVert+2\langle y_{n_{k}}-x_{n_{k}},\xi_{n_{k}}\rangle<0 for all kKk\geq K. Applying the triangle-inequality with respect to the points xnkx_{n_{k}} yields

ynkx2\displaystyle\left\lVert y_{n_{k}}-x\right\rVert^{2} =ε2+xnkx(xnkx+2ynkxnk,ξnk)<ε2\displaystyle=\varepsilon^{2}+\left\lVert x_{n_{k}}-x\right\rVert\left(\left\lVert x_{n_{k}}-x\right\rVert+2\langle y_{n_{k}}-x_{n_{k}},\xi_{n_{k}}\rangle\right)<\varepsilon^{2}

for all kKk\geq K. This implies the contradiction xint(Eε)x\in\textrm{int}(E_{\varepsilon}).

Step 3. ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}): For each nn\in\mathbb{N}, write rn:=xnxr_{n}:=\left\lVert x_{n}-x\right\rVert. Since (xn)n=1Eε(x_{n})_{n=1}^{\infty}\subset\partial E_{\varepsilon} there exists for each nn\in\mathbb{N} some znBrn2(xn)Eεcz_{n}\in B_{r_{n}^{2}}(x_{n})\cap E_{\varepsilon}^{c}. Then ξnz:=(znx)/znxξ\xi_{n}^{z}:=(z_{n}-x)/\left\lVert z_{n}-x\right\rVert\to\xi so that ξΞx(Eε)\xi\in\Xi_{x}(E_{\varepsilon}). Steps 1. and 2. together imply yx,ξ=0\langle y-x,\xi\rangle=0 so that ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) (see Definition 2.6). This concludes the proof of (ii)(a).

(ii)(b) Assume contrary to the claim that there exists some δ>0\delta>0 and a subsequence (ηnk)k=1(\eta_{n_{k}})_{k=1}^{\infty}, for which ηnk,ξ<1δ\left\lVert\langle\eta_{n_{k}},\xi\rangle\right\rVert<1-\delta for all kk\in\mathbb{N}. This implies that if ykΠEext(xnk)y_{k}\in\Pi_{E}^{\mathrm{ext}}(x_{n_{k}}) with ykxnk,ηnk=0\langle y_{k}-x_{n_{k}},\eta_{n_{k}}\rangle=0, there exists some r>0r>0, depending on δ\delta, for which

(2.5) ykxnk,ξr\left\lVert\langle y_{k}-x_{n_{k}},\xi\rangle\right\rVert\geq r

for infinitely many kk\in\mathbb{N}. On the other hand property (ii)(a) implies the existence of a subsequence (ykj)j=1(y_{k_{j}})_{j=1}^{\infty} for which the limit y:=limjykjΠEext(x)y:=\lim_{j\to\infty}y_{k_{j}}\in\Pi_{E}^{\mathrm{ext}}(x) exists and satisfies yx,ξ=0\langle y-x,\xi\rangle=0. Inequality (2.5) now leads to the contradiction

yx,ξ=limjykjxnkj,ξr>0.\left\lVert\langle y-x,\xi\rangle\right\rVert=\lim_{j\to\infty}\big{\|}\big{\langle}y_{k_{j}}-x_{n_{k_{j}}},\xi\big{\rangle}\big{\|}\geq r>0.

(i) According to (ii)(a) every convergent subsequence (ξnk)k=1(\xi_{n_{k}})_{k=1}^{\infty} satisfies ξnkξΞxext(Eε)\xi_{n_{k}}\to\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) as kk\to\infty. For Ξxext(Eε)={ξ}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi\} (a singleton) this implies ξnξ\xi_{n}\to\xi and the claim follows. In case Ξxext(Eε)={ξ(1),ξ(2)}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\left\{\xi^{(1)},\xi^{(2)}\right\} for some ξ(1)ξ(2)\xi^{(1)}\neq\xi^{(2)}, write r=ξ(1),ξ(2)r=\big{\|}\xi^{(1)},\xi^{(2)}\big{\|}. The compactness of Eε\partial E_{\varepsilon} implies that there exists some NN\in\mathbb{N}, for which

ξnBr/3(ξ(1))Br/3(ξ(2))\xi_{n}\in B_{r/3}\big{(}\xi^{(1)}\big{)}\cup B_{r/3}\big{(}\xi^{(2)}\big{)}

for all nNn\geq N. For each i{1,2}i\in\{1,2\}, define Ni:={n:ξnBr/3(ξ(i))}N_{i}:=\left\{n\in\mathbb{N}\,:\,\xi_{n}\in B_{r/3}\left(\xi^{(i)}\right)\right\}. If NiN_{i} is finite for some i{1,2}i\in\{1,2\}, we have limnξn=ξ(j)\lim_{n\to\infty}\xi_{n}=\xi^{(j)} for j{1,2}{i}j\in\{1,2\}\setminus\{i\} and the claim follows. Otherwise the sequences (ξ1,k)kN1(\xi_{1,k})_{k\in N_{1}} and (ξ2,k)kN1(\xi_{2,k})_{k\in\mathbb{N}\setminus N_{1}} are disjoint and satisfy limkξi,k=ξ(i)\lim_{k\to\infty}\xi_{i,k}=\xi^{(i)} for i{1,2}i\in\{1,2\}. ∎

Proposition 2.14 (Extremal outward directions coincide with tangents).

Let E2E\subset\mathbb{R}^{2} be compact and let xEεx\in\partial E_{\varepsilon}. Then Tx(Eε)=Ξxext(Eε)T_{x}(E_{\varepsilon})=\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}).

Proof.

Lemma 2.13 (ii)(a) implies Tx(Eε)Ξxext(Eε)T_{x}(E_{\varepsilon})\subset\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}), so we are left with proving the other direction. Assume ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}). Then there exists a sequence (xn)n=1Eεc(x_{n})_{n=1}^{\infty}\subset E_{\varepsilon}^{c}, for which

φnφn:=xnxxnxξ,\frac{\varphi_{n}}{\left\lVert\varphi_{n}\right\rVert}:=\frac{x_{n}-x}{\left\lVert x_{n}-x\right\rVert}\longrightarrow\xi,

as nn\to\infty. Since ξ\xi is an extremal outward direction, there exists some extremal contributor yΠEext(x)y\in\Pi_{E}^{\mathrm{ext}}(x) for which yx,ξ=0\langle y-x,\xi\rangle=0. Let y^:=(yx)/yx=(yx)/ε\widehat{y}:=(y-x)/\left\lVert y-x\right\rVert=(y-x)/\varepsilon and define Hn:=hnξ+rny^H_{n}:=h_{n}\xi+r_{n}\widehat{y}, where

hn:=φn1φn24ε2,andrn:=φn22ε.h_{n}:=\left\lVert\varphi_{n}\right\rVert\sqrt{1-\frac{\left\lVert\varphi_{n}\right\rVert^{2}}{4\varepsilon^{2}}},\quad\text{and}\quad r_{n}:=\frac{\left\lVert\varphi_{n}\right\rVert^{2}}{2\varepsilon}.

It follows from the orthogonality of ξ\xi and y^\widehat{y} that Hn=φn\left\lVert H_{n}\right\rVert=\left\lVert\varphi_{n}\right\rVert. Furthermore x+HnEεx+H_{n}\in E_{\varepsilon} for all nn\in\mathbb{N}, since (x+Hn)y=ε\left\lVert\left(x+H_{n}\right)-y\right\rVert=\varepsilon. See figure 3.

Consider now the φn\left\lVert\varphi_{n}\right\rVert-radius circle Bφn(x)\partial B_{\left\lVert\varphi_{n}\right\rVert}(x) centered at xx. The geodesic arc-segment (shortest path) on this circle that connects the points x+HnEεx+H_{n}\in E_{\varepsilon} and x+φn=xnEεcx+\varphi_{n}=x_{n}\in E_{\varepsilon}^{c} must necessarily contain a boundary point znBφn(x)Eεz_{n}\in\partial B_{\left\lVert\varphi_{n}\right\rVert}(x)\cap\partial E_{\varepsilon}.

Refer to caption
(a) The geometric picture. Here rn=εεcos(arcsin(hnε))=εε2hn2r_{n}=\varepsilon-\varepsilon\cos(\arcsin(\frac{h_{n}}{\varepsilon}))=\varepsilon-\sqrt{\varepsilon^{2}-h_{n}^{2}}, and one can solve for these values so that Hn=φn\left\lVert H_{n}\right\rVert=\left\lVert\varphi_{n}\right\rVert is satisfied.
Refer to caption
(b) For each nn\in\mathbb{N}, the point znEεz_{n}\in\partial E_{\varepsilon} lies on a geodesic arc-segment on Bφn(x)\partial B_{\left\lVert\varphi_{n}\right\rVert}(x), which connects the points xnx_{n} and x+Hnx+H_{n}.
Figure 3. The construction of the sequence (zn)n=1(z_{n})_{n=1}^{\infty}. Here the point xx is depicted as a wedge, but the procedure is the same for other types of boundary points.

Let δ>0\delta>0. Since xnxx_{n}\to x and φn/φnξ\varphi_{n}/\left\lVert\varphi_{n}\right\rVert\to\xi, there exists some NN\in\mathbb{N} for which

(2.6) φnδ3andφnφnξδ3\left\lVert\varphi_{n}\right\rVert\leq\frac{\delta}{3}\quad\text{and}\quad\left\lVert\frac{\varphi_{n}}{\left\lVert\varphi_{n}\right\rVert}-\xi\right\rVert\leq\frac{\delta}{3}

whenever nNn\geq N. Since

HnHnξ2\displaystyle\left\lVert\frac{H_{n}}{\left\lVert H_{n}\right\rVert}-\xi\right\rVert^{2} =24φn2ε20\displaystyle=2-\sqrt{4-\frac{\left\lVert\varphi_{n}\right\rVert^{2}}{\varepsilon^{2}}}\longrightarrow 0

as φn0\varphi_{n}\to 0, we can choose some NNN^{*}\geq N for which the inequality Hn/Hnξδ/3\left\lVert H_{n}/\left\lVert H_{n}\right\rVert-\xi\right\rVert\leq\delta/3 as well as the estimates (2.6) hold for all nNn\geq N^{*}. It follows from the definition of the points znz_{n} that zn(x+φn)φnHn\left\lVert z_{n}-(x+\varphi_{n})\right\rVert\leq\left\lVert\varphi_{n}-H_{n}\right\rVert for all nn\in\mathbb{N}. This allows us to obtain the estimate

znxznxξ\displaystyle\left\lVert\frac{z_{n}-x}{\left\lVert z_{n}-x\right\rVert}-\xi\right\rVert HnHnξ+2φnφnξδ,\displaystyle\leq\left\lVert\frac{H_{n}}{\left\lVert H_{n}\right\rVert}-\xi\right\rVert+2\left\lVert\frac{\varphi_{n}}{\left\lVert\varphi_{n}\right\rVert}-\xi\right\rVert\leq\delta,

which is valid for all nNn\geq N^{*}. Since also 0<znx=φn<δ0<\left\lVert z_{n}-x\right\rVert=\left\lVert\varphi_{n}\right\rVert<\delta for all nNn\geq N^{*}, we see that ξ\xi fulfils the requirements of Definition 2.3, so that ξTx(Eε)\xi\in T_{x}(E_{\varepsilon}). ∎

3. Local Structure of the Boundary

In this section we utilise the results obtained in Section 2 regarding outward directions and contributors in order to analyse the local properties of the boundary Eε\partial E_{\varepsilon}.

We begin by proving a local contribution property, Proposition 3.1, which intuitively states that in order to describe the local geometry of Eε\partial E_{\varepsilon} near a boundary point xEεx\in\partial E_{\varepsilon} it suffices to consider the geometry of E\partial E around the extremal contributors yΠEext(x)y\in\Pi_{E}^{\mathrm{ext}}(x).

In Section 3.2 we develop a method for approximating the set EεE_{\varepsilon} with finite collections of balls {Bε(dn):dnDn}\{B_{\varepsilon}(d_{n})\,:\,d_{n}\in D^{n}\} that correspond to certain finite subsets DnED^{n}\subset E. Combining this idea with Proposition 3.1 we proceed to show in Proposition 3.5 that local representations for the boundary Eε\partial E_{\varepsilon} may be obtained using finite collections of curves that can be represented as graphs of continuous functions on a compact interval.

As the first application of Proposition 3.5 we show in Lemma 3.8 that for every xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E) and every wedge (see Definition 4.1 and Figure 1) there exists a unique connected component VV of the complement EεcE_{\varepsilon}^{c} for which xVx\in\partial V.

Refer to caption
(a) The boundaries Bε(y1)\partial B_{\varepsilon}(y_{1}) and Bε(y2)\partial B_{\varepsilon}(y_{2}) give an approximation for the local geometry of the boundary Eε\partial E_{\varepsilon} inside the ball Br(x)B_{r}(x).
Refer to caption
(b) For an exact representation, one needs to consider all the contributors within some radius δ>0\delta>0 from the extremal contributors y1y_{1} and y2y_{2}.
Figure 4. Idea of local contribution. The points y1,y2y_{1},y_{2} are the extremal contributors of the wedge xEεx\in\partial E_{\varepsilon}.

3.1. Local Contribution

Intuitively, one can give a crude approximation for the boundary around each xEεx\in\partial E_{\varepsilon} by considering the boundaries Bε(y)\partial B_{\varepsilon}(y) centered at the contributors yΠE(x)y\in\Pi_{E}(x), and zooming in on a suitably small neighbourhood Br(x)B_{r}(x), in which

(3.1) EεBr(x)(yΠE(x)Bε(y))Br(x).\partial E_{\varepsilon}\cap B_{r}(x)\approx\partial\left(\bigcup_{y\in\Pi_{E}(x)}B_{\varepsilon}(y)\right)\cap B_{r}(x).

However, inside any neighbourhood Br(x)B_{r}(x) the geometry of the ε\varepsilon-neighbourhood EεE_{\varepsilon} is not defined solely by the positions of the contributors yΠE(x)y\in\Pi_{E}(x) (see Figures 4 and 5). Hence one needs to consider at least all the contributors in some neighbourhood of the set of extremal contributors ΠEext(x)\Pi_{E}^{\mathrm{ext}}(x). Proposition 3.1 below confirms that this is indeed sufficient.

We introduce here the following notation for open xx-centered half-balls oriented in the direction of some vS1v\in S^{1}:

(3.2) Ur(x,v):={zBr(x):zx,v>0}.U_{r}(x,v):=\{z\in B_{r}(x)\,:\,\langle z-x,v\rangle>0\}.
Proposition 3.1 (Local contribution).

Let E2E\subset\mathbb{R}^{2} and xEεx\in\partial E_{\varepsilon} with Ξxext(Eε)={ξ1,ξ2}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi_{1},\xi_{2}\}, where we allow ξ1=ξ2\xi_{1}=\xi_{2}. Then for all δ>0\delta>0 there exists some r>0r>0 such that given zBr(x)z\in B_{r}(x), we have zEεz\in E_{\varepsilon} if and only if either

(3.3) z\displaystyle z Ur(x,ξ1)¯Ur(x,ξ2)¯,or\displaystyle\notin\overline{U_{r}(x,\xi_{1})}\cup\overline{U_{r}(x,\xi_{2})},\mathrm{or}
(3.4) z\displaystyle z Bε(EBδ(ΠEext(x)))¯.\displaystyle\in\overline{B_{\varepsilon}\big{(}E\cap B_{\delta}(\Pi_{E}^{\mathrm{ext}}(x))\big{)}}.
Proof.

Assume to the contrary that there exists some δ>0\delta>0, for which the claim fails. This means that for all r>0r>0 there exists some zBr(x)z\in B_{r}(x), for which either

  1. (1)

    zEεz\in E_{\varepsilon} and both (3.3) and (3.4) fail, or

  2. (2)

    zEεz\notin E_{\varepsilon} and one of the conditions (3.3) or (3.4) holds true.

This implies that there exists a sequence (zk)k=1(z_{k})_{k=1}^{\infty} with zkB1/k(x)z_{k}\in B_{1/k}(x), for which either condition (1) or (2) holds true for z=zkz=z_{k} for infinitely many indices kk\in\mathbb{N}. A corresponding subsequence (zn)n=1(z_{n})_{n=1}^{\infty} then satisfies znxz_{n}\to x as nn\to\infty and either

  1. (a)

    for all nn\in\mathbb{N} znEεz_{n}\in E_{\varepsilon} while conditions (3.3) and (3.4) both fail for z=znz=z_{n}, or

  2. (b)

    for all nn\in\mathbb{N} znEεz_{n}\notin E_{\varepsilon} while either condition (3.3) or (3.4) holds true for z=znz=z_{n}.

We proceed by showing that both of these statements lead to a contradiction. In both cases one can assume, without loss of generality, the existence of the limit

vz:=limn(znx)/znx.v_{z}:=\lim_{n\to\infty}(z_{n}-x)/\left\lVert z_{n}-x\right\rVert.
Refer to caption
Figure 5. Schematic illustration of Proposition 3.1. For a sufficiently small r>0r>0, the balls Bδ(y1)B_{\delta}(y_{1}) and Bδ(y2)B_{\delta}(y_{2}) contain all the contributors yΠE(z)y\in\Pi_{E}(z) (red dotted lines) of those boundary points zEεz\in\partial E_{\varepsilon} that lie inside the ball Br(x)B_{r}(x) (red solid lines). For ρ>r\rho>r the boundary segment generated by the point yEy^{*}\in E would lie inside the larger ball Bρ(x)B_{\rho}(x) and would need to be accounted for separately. Note that the geometry of the boundary Eε\partial E_{\varepsilon} inside Br(x)B_{r}(x) is not affected by whether or not the points yy on the blue dotted line satisfy yEy\in E.

Assume first that (a) holds true. This means that

(3.5) znEε(U1/n(x,ξ1)¯U1/n(x,ξ2)¯)andznBε(EBδ(ΠEext(x)))¯z_{n}\in E_{\varepsilon}\cap\left(\overline{U_{1/n}(x,\xi_{1})}\cup\overline{U_{1/n}(x,\xi_{2})}\right)\quad\mathrm{and}\quad z_{n}\notin\overline{B_{\varepsilon}\big{(}E\cap B_{\delta}(\Pi_{E}^{\mathrm{ext}}(x))\big{)}}

for all nn\in\mathbb{N}. Since znEεz_{n}\in E_{\varepsilon} for all nn\in\mathbb{N}, one can choose a sequence (yn)n=1E(y_{n})_{n=1}^{\infty}\subset E with znBε(yn)¯z_{n}\in\overline{B_{\varepsilon}(y_{n})}. In addition, since znxz_{n}\to x, Lemma 2.8 guarantees the existence of a convergent subsequence (ynk)k=1(y_{n_{k}})_{k=1}^{\infty} with the limit y^:=limkynkΠE(x)\widehat{y}:=\lim_{k\to\infty}y_{n_{k}}\in\Pi_{E}(x). Note that it follows from (3.5) that ynBδ(ΠEext(x))y_{n}\notin B_{\delta}(\Pi_{E}^{\mathrm{ext}}(x)) for all nn\in\mathbb{N}, which implies y^ΠEext(x)\widehat{y}\notin\Pi_{E}^{\mathrm{ext}}(x), and consequently ΠE(x)ΠEext(x)\Pi_{E}(x)\setminus\Pi_{E}^{\mathrm{ext}}(x)\neq\varnothing. This rules out the possibility ξ1=ξ2\xi_{1}=-\xi_{2} for the extremal outward directions ξ1,ξ2Ξxext(Eε)\xi_{1},\xi_{2}\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}), and therefore

(3.6) ξ1,ξ2>1.\langle\xi_{1},\xi_{2}\rangle>-1.

The relations (3.5) guarantee that znkBε(ΠEext(x))z_{n_{k}}\notin B_{\varepsilon}(\Pi_{E}^{\mathrm{ext}}(x)), which together with Lemma 2.9 (i) implies yx,vz0\langle y-x,v_{z}\rangle\leq 0 for all extremal contributors yΠEext(x)y\in\Pi_{E}^{\mathrm{ext}}(x). Combined with the assumption that

znkU1/k(x,ξ1)¯U1/k(x,ξ2)¯z_{n_{k}}\in\overline{U_{1/k}(x,\xi_{1})}\cup\overline{U_{1/k}(x,\xi_{2})}

for all kk\in\mathbb{N}, this implies that vz[ξ1,ξ2]S1=Ξx(Eε)v_{z}\in[\xi_{1},\xi_{2}]_{S^{1}}=\Xi_{x}(E_{\varepsilon}) (see Proposition 2.12), where [ξ1,ξ2]S1[\xi_{1},\xi_{2}]_{S^{1}} is a geodesic arc-segment due to (3.6). It follows that y^x,vz<0\langle\widehat{y}-x,v_{z}\rangle<0, since y^ΠEext(x)\widehat{y}\notin\Pi_{E}^{\mathrm{ext}}(x). But now a computation analogous to that presented in the proof of Lemma 2.9 (ii) leads to the contradiction y^x<ε\left\lVert\widehat{y}-x\right\rVert<\varepsilon.

Assume then that (b) holds true. Now znEεz_{n}\notin E_{\varepsilon} for all nn\in\mathbb{N} so that vzΞx(Eε)v_{z}\in\Xi_{x}(E_{\varepsilon}). In addition, given that either (3.3) or (3.4) is satisfied for each znz_{n} and (3.4) implies znEεz_{n}\in E_{\varepsilon}, condition (3.3) necessarily holds true for all znz_{n}. Then ξ1ξ2\xi_{1}\neq-\xi_{2}, since ξ1=ξ2\xi_{1}=-\xi_{2} leads to the contradiction

znB1/n(x)i{1,2}U1/n(x,ξi)¯=B1/n(x)B1/n(x)=.z_{n}\in B_{1/n}(x)\setminus\bigcup_{i\in\{1,2\}}\overline{U_{1/n}(x,\xi_{i})}=B_{1/n}(x)\setminus B_{1/n}(x)=\varnothing.

On the other hand, if ξ1ξ2\xi_{1}\neq-\xi_{2}, Proposition 2.12 states that vzΞx(Eε)v_{z}\in\Xi_{x}(E_{\varepsilon}) can be written as a convex combination vz=aξ1+bξ2v_{z}=a\xi_{1}+b\xi_{2}. Note that at least one of the coefficients a,ba,b must be strictly positive, since vzS1v_{z}\in S^{1}. However, (3.3) implies znx,ξi<0\langle z_{n}-x,\xi_{i}\rangle<0 for i{1,2}i\in\{1,2\} and all nn\in\mathbb{N}, which leads to the contradiction vz,ξi0\langle v_{z},\xi_{i}\rangle\leq 0 for i{1,2}i\in\{1,2\}. ∎

3.2. Approximating the Boundary with Continuous Graphs

In order to study the properties of the boundary Eε\partial E_{\varepsilon}, we develop a finite approximation scheme as a technical aid. The idea is to generate an expanding sequence (Dn)n=1(D^{n})_{n=1}^{\infty} of finite subsets DnED^{n}\subset E and consider their ε\varepsilon-neighbourhoods Bε(Dn)B_{\varepsilon}(D^{n}), whose boundaries approximate the actual boundary Eε\partial E_{\varepsilon} uniformly with respect to Hausdorff distance.

Definition 3.2 (Hausdorff distance).

The Hausdorff distance between X,YdX,Y\subset\mathbb{R}^{d} is

distH(X,Y):=inf{δ>0:XYδandYXδ},\mathrm{dist}_{H}(X,Y):=\mathrm{inf}\big{\{}\delta>0\,:\,X\subset Y_{\delta}\,\,\textrm{and}\,\,Y\subset X_{\delta}\big{\}},

where Xδ:=xXBδ(x)¯X_{\delta}:=\overline{\bigcup_{x\in X}B_{\delta}(x)}.

Let E2E\subset\mathbb{R}^{2} be compact, ε>0\varepsilon>0 and n0n_{0}\in\mathbb{N} with n0>ln(4/ε)ln2n_{0}>\frac{\ln(4/\varepsilon)}{\ln 2}. Consider for natural numbers n>n0n>n_{0} the partitions of 2\mathbb{R}^{2} into squares

𝒞n:={Ck,n:=[k2n,k+12n]×[2n,+12n]:k,}.\mathcal{C}^{n}:=\left\{C_{k,\ell}^{n}:=\left[\frac{k}{2^{n}},\frac{k+1}{2^{n}}\right]\times\left[\frac{\ell}{2^{n}},\frac{\ell+1}{2^{n}}\right]:k,\ell\in\mathbb{Z}\right\}.

Due to the Axiom of Choice, there exists a non-decreasing sequence Dn0Dn0+1D^{n_{0}}\subset D^{n_{0}+1}\subset\dots of subsets of DnD^{n} of 2\mathbb{R}^{2} such that

DnCk,n={{dk,n}ifECk,n,ifECk,n=,D^{n}\cap C_{k,\ell}^{n}=\left\{\begin{array}[]{c@{\quad\text{if}\,\,}l}\big{\{}d_{k,\ell}^{n}\big{\}}&E\cap C_{k,\ell}^{n}\not=\emptyset,\\ \emptyset&E\cap C_{k,\ell}^{n}=\emptyset,\end{array}\right.

where the points dk,nd_{k,\ell}^{n} are chosen arbitrarily from ECk,nE\cap C_{k,\ell}^{n}. It is easy to verify that Eε=nn0Bε(Dn)E_{\varepsilon}=\bigcup_{n\geq n_{0}}B_{\varepsilon}(D^{n}) and that the approximations converge to EεE_{\varepsilon} in Hausdorff distance,

(3.7) limndistH(Eε,Bε(Dn))=0.\lim_{n\to\infty}\operatorname{dist}_{H}\big{(}E_{\varepsilon},B_{\varepsilon}(D^{n})\big{)}=0\,.
Refer to caption
(a) The boundaries Bε(dk,n)\partial B_{\varepsilon}(d_{k,\ell}^{n}) for dk,nDnd_{k,\ell}^{n}\in D^{n} (red dots) provide an approximation (red curve) for Eε\partial E_{\varepsilon}.
Refer to caption
(b) For Dn+1D^{n+1} the approximation improves, and as nn\to\infty, it converges uniformly to Eε\partial E_{\varepsilon}.
Figure 6. A schematic illustration of two consecutive finite approximating sets DnD^{n} and Dn+1D^{n+1} near a wedge xx.
Definition 3.3 (Finite approximating sets).

Let E2E\subset\mathbb{R}^{2} be compact and let (Dn)n=1(D^{n})_{n=1}^{\infty} be a sequence of subsets DnED^{n}\subset E as described above. Then the sets DnD^{n} are called finite approximating sets for the set EE.

We now have the necessary ingredients in place for defining what we call local boundary representations near each boundary point xEεx\in\partial E_{\varepsilon}. The local contribution property, Proposition 3.1, implies that in order to describe the boundary Eε\partial E_{\varepsilon} near xx, one only needs to consider contributors yEy\in E around the extremal contributors yΠEext(x)y\in\Pi_{E}^{\mathrm{ext}}(x). The extremal outward directions ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) together with their corresponding extremal contributors form extremal pairs (ξ,y)(\xi,y) that represent coordinate axes adapted to the orientation of the boundary near xx.

Definition 3.4 (Extremal pairs).

Let EdE\subset\mathbb{R}^{d} be closed, let xEεx\in\partial E_{\varepsilon} and denote by Ξxext(Eε)\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) and ΠEext(x)\Pi_{E}^{\mathrm{ext}}(x) the sets of extremal outward directions and extremal contributors, respectively. We define the set of extremal pairs at xx as the collection

𝒫xext(Eε):={(ξ,y)Ξxext(Eε)×ΠEext(x):yx,ξ=0}.\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}):=\big{\{}(\xi,y)\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})\times\Pi_{E}^{\mathrm{ext}}(x)\,:\,\langle y-x,\xi\rangle=0\big{\}}.

One can thus choose a finite approximating sequence DnED^{n}\subset E, interpret the boundaries Bε(Dn)\partial B_{\varepsilon}(D^{n}) near xx as graphs of continuous functions fnf_{n} in the coordinate system (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) and obtain, as a uniform limit, a continuous function f=limnfnf=\lim_{n\to\infty}f_{n} whose graph serves as a representation of a part of the boundary Eε\partial E_{\varepsilon} near xx. Using this construction for each extremal pair (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) one obtains a finite set of continuous graphs that give a complete representation of the boundary near xx.

We note that the proof of Proposition 3.5 below makes use of Lemma 3.7 which we have placed in the subsequent Section 3.3, together with other results on the geometry of the complement EεcE_{\varepsilon}^{c}.

Proposition 3.5 (Local boundary representation).

Let E2E\subset\mathbb{R}^{2} be closed and let xEεx\in\partial E_{\varepsilon}. For each extremal pair (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) there exists a continuous function fξ,y:[0,ε/2]f^{\xi,y}:[0,\varepsilon/2]\to\mathbb{R} and a corresponding function gξ,y:[0,ε/2]2g_{\xi,y}:[0,\varepsilon/2]\to\mathbb{R}^{2}, given by

(3.8) gξ,y(s):=x+sξ+fξ,y(s)(xy),g_{\xi,y}(s):=x+s\xi+f^{\xi,y}(s)(x-y),

so that the collection 𝒢(x):={gξ,y:(ξ,y)𝒫xext(Eε)}\mathcal{G}(x):=\left\{g_{\xi,y}\,:\,(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon})\right\} satisfies

(3.9) EεBr(x)¯=(ξ,y)𝒫xext(Eε)gξ,y(Aξ,y)\partial E_{\varepsilon}\cap\overline{B_{r}(x)}=\bigcup_{(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon})}g_{\xi,y}\left(A_{\xi,y}\right)

for some r>0r>0 and some closed Aξ,y[0,ε/2]A_{\xi,y}\subset[0,\varepsilon/2]. We call the collection 𝒢(x)\mathcal{G}(x) a local boundary representation (of radius rr) at xx. For each extremal pair (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) the corresponding subset Aξ,y[0,ε/2]A_{\xi,y}\subset[0,\varepsilon/2] is either

  1. (a)

    an interval [0,sξ,y][0,s_{\xi,y}] for some 0<sξ,yε/20<s_{\xi,y}\leq\varepsilon/2, or

  2. (b)

    a closed set whose complement in [0,ε/2][0,\varepsilon/2] contains a sequence of disjoint open intervals with 0 as an accumulation point.

For wedges (type S1) and xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E), case (a) above holds true for all (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}).

Proof.

Define for each extremal pair (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) the sets

Cξ,y\displaystyle C_{\xi,y} :={x+sξ+t(xy):(s,t)[ε,ε]×(,ε/2]},\displaystyle:=\left\{x+s\xi+t(x-y)\,:\,(s,t)\in[-\varepsilon,\varepsilon]\times(-\infty,-\varepsilon/2]\right\},
Eξ,y\displaystyle E_{\xi,y} :=ECξ,y.\displaystyle:=E\cap C_{\xi,y}.

Consider a sequence (Dn)n=1(D_{n})_{n=1}^{\infty} of finite approximating sets for the set EE (see Definition 3.3), and define Dnξ,y:=Eξ,yDnD_{n}^{\xi,y}:=E_{\xi,y}\cap D_{n} for each nn\in\mathbb{N}. Using the sets Dnξ,yD_{n}^{\xi,y} we define a sequence of functions fnξ,y:[ε/2,ε/2]f_{n}^{\xi,y}:[-\varepsilon/2,\varepsilon/2]\to\mathbb{R} by setting

(3.10) fnξ,y(s):=max{t:dist(x+sξ+t(xy),Dnξ,y)ε},f_{n}^{\xi,y}(s):=\textrm{max}\left\{t\in\mathbb{R}\,:\,\mathrm{dist}\left(x+s\xi+t(x-y),D_{n}^{\xi,y}\right)\leq\varepsilon\right\},

where dist(,Dnξ,y)\textrm{dist}\left(\cdot,D_{n}^{\xi,y}\right) denotes the Euclidean distance from the set Dnξ,yD_{n}^{\xi,y}. The maximum in (3.10) exists for all s[ε/2,ε/2]s\in[-\varepsilon/2,\varepsilon/2] due to the compactness of Bε(Dnξ,y)¯\overline{B_{\varepsilon}\big{(}D_{n}^{\xi,y}\big{)}}, which also implies that the functions fnξ,yf_{n}^{\xi,y} are bounded for all nn\in\mathbb{N}. In addition, it follows from the definition of the values fnξ,y(s)f_{n}^{\xi,y}(s) that

x+sξ+fnξ,y(s)(xy)Bε(Dnξ,y)¯x+s\xi+f_{n}^{\xi,y}(s)(x-y)\in\partial\overline{B_{\varepsilon}\big{(}D_{n}^{\xi,y}\big{)}}

for all s[ε/2,ε/2]s\in[-\varepsilon/2,\varepsilon/2]. Since for each nn\in\mathbb{N} the boundary Bε(Dnξ,y)¯\partial\overline{B_{\varepsilon}\big{(}D_{n}^{\xi,y}\big{)}} is composed of a finite collection of arc-segments, the corresponding functions fnξ,yf_{n}^{\xi,y} are continuous.

It is easy to verify that the boundaries Bε(Dn)¯\partial\overline{B_{\varepsilon}\left(D_{n}\right)} converge to the boundary Eε\partial E_{\varepsilon} uniformly in Hausdorff distance. Similarly the boundaries Bε(Dnξ,y)¯\partial\overline{B_{\varepsilon}\big{(}D_{n}^{\xi,y}\big{)}} converge uniformly to the boundaries Bε(Eξ,y)¯\partial\overline{B_{\varepsilon}(E_{\xi,y})}. Hence (fnξ,y)n=1\left(f_{n}^{\xi,y}\right)_{n=1}^{\infty} is a uniformly convergent, monotonically increasing sequence of continuous functions on the compact interval [ε/2,ε/2][-\varepsilon/2,\varepsilon/2], which implies that the limiting function

(3.11) fξ,y(s):=limnfnξ,y(s)f^{\xi,y}(s):=\lim_{n\to\infty}f_{n}^{\xi,y}(s)

is continuous. One can therefore define for each (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) a function gξ,y:[0,ε/2]2g_{\xi,y}:[0,\varepsilon/2]\to\mathbb{R}^{2} by

(3.12) gξ,y(s):=x+sξ+fξ,y(s)(xy).g_{\xi,y}(s):=x+s\xi+f^{\xi,y}(s)(x-y).

The remainder of the proof concerns the analysis of the relationship between the graphs gξ,y([0,ε/2])g_{\xi,y}([0,\varepsilon/2]) and the boundary Eε\partial E_{\varepsilon} near the point xx, with the aim of identifying subsets Aξ,y[0,ε/2]A_{\xi,y}\subset[0,\varepsilon/2] that satisfy (3.9).

Refer to caption
(a) n=1n=1
Refer to caption
(b) n=2n=2
Figure 7. First functions in the sequence (fnξ,y)n=1\left(f_{n}^{\xi,y}\right)_{n=1}^{\infty}. The red dots represent the finite approximating sets D1ξ,yD_{1}^{\xi,y} and D2ξ,yD_{2}^{\xi,y}.

Substituting δ=ε/2\delta=\varepsilon/2 in Proposition 3.1 guarantees the existence of some r>0r>0, for which zEεBr(x)z\in E_{\varepsilon}\cap B_{r}(x) if and only if either

(i) zUr(x,ξ)¯ for each ξΞxext(Eε),or(ii) zBε(EBε/2(ΠEext(x)))¯.\textrm{(i) }z\notin\overline{U_{r}(x,\xi)}\textrm{ for each }\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}),\quad\textrm{or}\quad\textrm{(ii) }z\in\overline{B_{\varepsilon}\left(E\cap B_{\varepsilon/2}(\Pi_{E}^{\mathrm{ext}}(x))\right)}.

According to Proposition 2.14, the set of tangential directions Tx(Eε)T_{x}(E_{\varepsilon}) coincides with the set of extremal outward directions Ξxext(Eε)\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}). This implies that condition (i) necessarily fails for all boundary points zEεBr(x)z\in\partial E_{\varepsilon}\cap B_{r}(x). Therefore condition (ii) holds true whenever zEεBr(x)z\in\partial E_{\varepsilon}\cap B_{r}(x), in which case zBε(Eξ,y)¯z\in\partial\overline{B_{\varepsilon}(E_{\xi,y})} for some (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}). Hence there exists some sz[0,ε/2]s_{z}\in[0,\varepsilon/2], for which z=gξ,y(sz)z=g_{\xi,y}(s_{z}). In other words, every boundary point zEεz\in\partial E_{\varepsilon} inside the neighbourhood Br(x)B_{r}(x) lies on one of the graphs gξ,y([0,ε/2])g_{\xi,y}([0,\varepsilon/2]) with (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}).

It remains to be verified, for each (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}), which arguments s[0,ε/2]s\in[0,\varepsilon/2] satisfy gξ,y(s)EεBr(x)g_{\xi,y}(s)\in\partial E_{\varepsilon}\cap B_{r}(x). We divide the rest of the proof into two cases, depending on whether or not there exist extremal contributors y1,y2ΠEext(x)y_{1},y_{2}\in\Pi_{E}^{\mathrm{ext}}(x) for which

(3.13) y1x=(y2x).y_{1}-x=-(y_{2}-x).

Condition (3.13) is schematically illustrated in Figure 9 (c) below.

(i) In case (3.13) is not satisfied by any y1,y2ΠEext(x)y_{1},y_{2}\in\Pi_{E}^{\mathrm{ext}}(x), there is either only one contributor so that xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E), or else xx is a wedge (see Definition 4.1). In both cases we have 𝒫xext(Eε)={(ξ1,y1),(ξ2,y2)}\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{(\xi_{1},y_{1}),(\xi_{2},y_{2})\} for some ξ1,ξ2Ξext(x)\xi_{1},\xi_{2}\in\Xi_{\textrm{ext}}(x) and y1,y2ΠEext(x)y_{1},y_{2}\in\Pi_{E}^{\mathrm{ext}}(x), where we allow for y1=y2y_{1}=y_{2} in case xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E). It follows from Lemma 3.7 that for any outward directions η1,η2(ξ1,ξ2)S1\eta_{1},\eta_{2}\in(\xi_{1},\xi_{2})_{S^{1}} there exists a neighbourhood Bρ(x)Br(x)B_{\rho}(x)\subset B_{r}(x) in which the graphs of gξ1,y1g_{\xi_{1},y_{1}} and gξ2,y2g_{\xi_{2},y_{2}} are separated by the cone Vρ(x,η1,η2)V_{\rho}(x,\eta_{1},\eta_{2}). Hence

(3.14) gξ1,y1([0,ε/2])gξ2,y2([0,ε/2])Bρ(x)=.g_{\xi_{1},y_{1}}([0,\varepsilon/2])\cap g_{\xi_{2},y_{2}}([0,\varepsilon/2])\cap B_{\rho}(x)=\varnothing.

For i{1,2}i\in\{1,2\} we define the upper bounds

sξi,yi:=max{s[0,ε/2]:gξi,yi([0,s])Bρ(x)¯}s_{\xi_{i},y_{i}}:=\max\Big{\{}s\in[0,\varepsilon/2]\,:\,g_{\xi_{i},y_{i}}\left([0,s]\right)\in\overline{B_{\rho}(x)}\Big{\}}

and the corresponding sets Aξi,yi:=[0,sξi,yi]A_{\xi_{i},y_{i}}:=\left[0,s_{\xi_{i},y_{i}}\right]. It follows then from (3.14) and Proposition 3.1 that gξi,yi(Aξi,yi)Eεg_{\xi_{i},y_{i}}(A_{\xi_{i},y_{i}})\subset\partial E_{\varepsilon} for i{1,2}i\in\{1,2\}, which implies

EεBρ(x)¯=i{1,2}gξi,yi(Aξi,yi).\partial E_{\varepsilon}\cap\overline{B_{\rho}(x)}=\bigcup_{i\in\{1,2\}}g_{\xi_{i},y_{i}}\left(A_{\xi_{i},y_{i}}\right).

(ii) Assume then that (3.13) holds true for y1,y2ΠEext(x)y_{1},y_{2}\in\Pi_{E}^{\mathrm{ext}}(x) and consider some ξΠEext(x)\xi\in\Pi_{E}^{\mathrm{ext}}(x). In this case the graphs gξ,yi([0,s])g_{\xi,y_{i}}([0,s]), i{1,2}i\in\{1,2\} may generally intersect for arbitrarily small s(0,ε/2]s\in(0,\varepsilon/2]. Due to (3.13) one can write for {i,j}={1,2}\{i,j\}=\{1,2\}

gξ,yi(s)\displaystyle g_{\xi,y_{i}}(s) =x+sξ+(fξ,yj(s)αξ(s))(xyj)\displaystyle=x+s\xi+\left(f^{\xi,y_{j}}(s)-\alpha_{\xi}(s)\right)(x-y_{j})
=gξ,yj(s)αξ(s)(xyj),\displaystyle=g_{\xi,y_{j}}(s)-\alpha_{\xi}(s)(x-y_{j}),

where αξ(s):=fξ,y1(s)+fξ,y2(s)\alpha_{\xi}(s):=f^{\xi,y_{1}}(s)+f^{\xi,y_{2}}(s). Hence it follows for i{1,2}i\in\{1,2\} from Proposition 3.1 and the definitions of the functions fnξ,yf_{n}^{\xi,y} and fξ,yf^{\xi,y} (see (3.10) and (3.11)) that

gξ,yi(s)\displaystyle g_{\xi,y_{i}}(s) Eεwheneverαξ(s)>0,and\displaystyle\notin\partial E_{\varepsilon}\,\,\mathrm{whenever}\,\,\alpha_{\xi}(s)>0,\,\mathrm{and}
gξ,yi(s)\displaystyle g_{\xi,y_{i}}(s) Eεwheneverαξ(s)<0.\displaystyle\in\partial E_{\varepsilon}\,\,\mathrm{whenever}\,\,\alpha_{\xi}(s)<0.

For αξ(s)=0\alpha_{\xi}(s)=0 one has gξ,y1(s)=gξ,y2(s)Eεg_{\xi,y_{1}}(s)=g_{\xi,y_{2}}(s)\in\partial E_{\varepsilon} if and only if there exists a sequence (sn)n=1(s_{n})_{n=1}^{\infty} with snss_{n}\to s, for which αξ(sn)<0\alpha_{\xi}(s_{n})<0 for all nn\in\mathbb{N}. This follows from the fact that

τgξ,y1(s)+(1τ)gξ,y2(s)Eεc\tau g_{\xi,y_{1}}(s)+(1-\tau)g_{\xi,y_{2}}(s)\in E_{\varepsilon}^{c}

whenever s(0,r)s\in(0,r), αξ(s)<0\alpha_{\xi}(s)<0 and τ(0,1)\tau\in(0,1). Equation (3.9) is therefore satisfied for the neighbourhood Br(x)B_{r}(x) and the sets

Aξ,yi:=Aξ:={s[0,r]:αξ(s)<0}¯={s[0,r]:gξ,y1(s),gξ,y2(s)Eε},A_{\xi,y_{i}}:=A_{\xi}:=\overline{\left\{s\in[0,r]\,:\,\alpha_{\xi}(s)<0\right\}}=\left\{s\in[0,r]\,:\,g_{\xi,y_{1}}(s),g_{\xi,y_{2}}(s)\in\partial E_{\varepsilon}\right\},

where i{1,2}i\in\{1,2\} and ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}). Now either [0,sξ]Aξ[0,s_{\xi}]\subset A_{\xi} for some sξ(0,r]s_{\xi}\in(0,r] or otherwise [0,T]Aξ[0,T]\setminus A_{\xi}\neq\varnothing for all T>0T>0. In any case 0 is an accumulation point of AξA_{\xi}, since xEεx\in\partial E_{\varepsilon} and ξTx(Eε)\xi\in T_{x}(E_{\varepsilon}) (see Proposition 2.14). ∎

Consider a boundary point xEεx\in\partial E_{\varepsilon} and the corresponding local boundary representation 𝒢(x)\mathcal{G}(x) with radius r>0r>0, and let zEεBr(x)z\in\partial E_{\varepsilon}\cap B_{r}(x) with

z=gξ,y(sz)=x+szξ+fξ,y(sz)(xy)z=g_{\xi,y}(s_{z})=x+s_{z}\xi+f^{\xi,y}(s_{z})(x-y)

for some (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) and sz[0,r]s_{z}\in[0,r]. The construction given in Proposition 3.5 guarantees that, when written in the (ξ,y)(\xi,y)-coordinates, the ξ\xi-coordinate sws_{w} of each contributor wΠE(z)w\in\Pi_{E}(z) satisfies sw[0,ε/2]s_{w}\in[0,\varepsilon/2]. This implies a lower and upper bound for the local growth-rate of the function fξ,yf^{\xi,y}. Combining this observation with Proposition 2.14 allows us to deduce that the functions fξ,yf^{\xi,y} in (3.8) are in fact Lipschitz continuous on [0,r][0,r] for all (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}), with a Lipschitz constant K=1/3εK=1/\sqrt{3}\varepsilon.

Proposition 3.6 (Local boundary representation is Lipschitz).

Let E2E\subset\mathbb{R}^{2}, let xEεx\in\partial E_{\varepsilon} and let 𝒢(x)\mathcal{G}(x) be a local boundary representation at xx with radius r>0r>0. For each extremal pair (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}), the function fξ,yf^{\xi,y} in (3.8) is 1/3ε1/\sqrt{3}\varepsilon-Lipschitz, and the function gξ,y𝒢(x)g_{\xi,y}\in\mathcal{G}(x) is 2/32/\sqrt{3}-Lipschitz on the interval [0,r][0,r].

Proof.

In order to work in an orthonormal coordinate system, we define for all (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) the functions hξ,y:[0,r]h^{\xi,y}:[0,r]\to\mathbb{R} by setting hξ,y(s):=εfξ,y(s)h^{\xi,y}(s):=\varepsilon f^{\xi,y}(s) so that

gξ,y(s)=x+sξ+hξ,y(s)ε1(xy)g_{\xi,y}(s)=x+s\xi+h^{\xi,y}(s)\varepsilon^{-1}(x-y)

for all gξ,y𝒢(x)g_{\xi,y}\in\mathcal{G}(x). We will show that the functions hξ,yh^{\xi,y} are 1/31/\sqrt{3}-Lipschitz, from which the claim follows.

Assume contrary to the claim that for some (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) there exist some p>1/3p>1/\sqrt{3} and s,w[0,r]s,w\in[0,r] for which |hξ,y(s)hξ,y(w)|p|sw||h^{\xi,y}(s)-h^{\xi,y}(w)|\geq p|s-w|. We present the argument for the case w<sw<s and hξ,y(s)hξ,y(w)p(sw)h^{\xi,y}(s)-h^{\xi,y}(w)\geq p(s-w). Assuming hξ,y(s)hξ,y(w)p(sw)h^{\xi,y}(s)-h^{\xi,y}(w)\leq-p(s-w) leads to a contradiction through similar reasoning. Write m:=(s+w)/2m:=(s+w)/2 and define

(w1,s1):={(m,s)ifhξ,y(s)hξ,y(m)>p(sm),(w,m)ifhξ,y(s)hξ,y(m)p(sm)(w_{1},s_{1}):=\left\{\begin{array}[]{c@{\quad\text{if}\,\,}l}(m,s)&h^{\xi,y}(s)-h^{\xi,y}(m)>p(s-m),\\ (w,m)&h^{\xi,y}(s)-h^{\xi,y}(m)\leq p(s-m)\end{array}\right.

so that hξ,y(s1)hξ,y(w1)p(s1w1)h^{\xi,y}(s_{1})-h^{\xi,y}(w_{1})\geq p(s_{1}-w_{1}). For nn\in\mathbb{N} we define inductively mn:=(sn+wn)/2m_{n}:=(s_{n}+w_{n})/2 and

(wn+1,sn+1):={(mn,sn)ifhξ,y(sn)hξ,y(mn)>p(snmn),(wn,mn)ifhξ,y(sn)hξ,y(mn)p(snmn)(w_{n+1},s_{n+1}):=\left\{\begin{array}[]{c@{\quad\text{if}\,\,}l}(m_{n},s_{n})&h^{\xi,y}(s_{n})-h^{\xi,y}(m_{n})>p(s_{n}-m_{n}),\\ (w_{n},m_{n})&h^{\xi,y}(s_{n})-h^{\xi,y}(m_{n})\leq p(s_{n}-m_{n})\end{array}\right.

so that for all nn\in\mathbb{N}

(3.15) hξ,y(sn)hξ,y(wn)p(snwn).h^{\xi,y}(s_{n})-h^{\xi,y}(w_{n})\geq p(s_{n}-w_{n}).

By construction, the sequence (sn)n=1(s_{n})_{n=1}^{\infty} is non-increasing, and the sequence (wn)n=1(w_{n})_{n=1}^{\infty} non-decreasing. Since |snwn|=2n|sw|0|s_{n}-w_{n}|=2^{-n}|s-w|\to 0 as nn\to\infty and sn>wns_{n}>w_{n} for all nn\in\mathbb{N}, there exists a unique limit a=limnsn=limnwna=\lim_{n\to\infty}s_{n}=\lim_{n\to\infty}w_{n}. From (3.15) it follows that

(3.16) pDn:=ks(n)hξ,y(sn)hξ,y(a)sna+kr(n)hξ,y(a)hξ,y(wn)awn,p\leq D_{n}:=k_{s}(n)\frac{h^{\xi,y}(s_{n})-h^{\xi,y}(a)}{s_{n}-a}+k_{r}(n)\frac{h^{\xi,y}(a)-h^{\xi,y}(w_{n})}{a-w_{n}},

where the coefficients

ks(n):=snasnwn,kw(n):=awnsnwnk_{s}(n):=\frac{s_{n}-a}{s_{n}-w_{n}},\quad k_{w}(n):=\frac{a-w_{n}}{s_{n}-w_{n}}

satisfy ks(n)+kw(n)=1k_{s}(n)+k_{w}(n)=1 for all nn\in\mathbb{N}. In case sN=as_{N}=a (resp. wN=aw_{N}=a) for some NN\in\mathbb{N}, the coefficient ks(n)k_{s}(n) (resp. kw(n)k_{w}(n)) vanishes for all nNn\geq N. Hence at least one and possibly both of the right and left derivatives of hξ,yh^{\xi,y} at aa are given by

D+hξ,y(a):=limnhξ,y(sn)hξ,y(a)sna,Dhξ,y(a):=limnhξ,y(a)hξ,y(wn)awnD^{+}h^{\xi,y}(a):=\lim_{n\to\infty}\frac{h^{\xi,y}(s_{n})-h^{\xi,y}(a)}{s_{n}-a},\quad D^{-}h^{\xi,y}(a):=\lim_{n\to\infty}\frac{h^{\xi,y}(a)-h^{\xi,y}(w_{n})}{a-w_{n}}

and correspond to the tangential directions on Eε\partial E_{\varepsilon} at x(a):=x+aξ+hξ,y(a)ε1(xy)x(a):=x+a\xi+h^{\xi,y}(a)\varepsilon^{-1}(x-y). According to Proposition 2.14 these in turn coincide with the extremal outward directions ξa+\xi_{a}^{+} and ξa\xi_{a}^{-} at x(a)x(a). Since the ξ\xi-coordinates of the corresponding extremal contributors ya+,yay_{a}^{+},y_{a}^{-} lie on the interval [0,ε/2][0,\varepsilon/2], the directional derivatives necessarily satisfy the upper bound (see Figure LABEL:Figure_Maximal_Tangent)

D±hξ,y(a)ε2ε2(ε2)2=13.D^{\pm}h^{\xi,y}(a)\leq\frac{\frac{\varepsilon}{2}}{\sqrt{\varepsilon^{2}-\big{(}\frac{\varepsilon}{2}\big{)}^{2}}}=\frac{1}{\sqrt{3}}.

The contradiction plimnDn=1/3<pp\leq\lim_{n\to\infty}D_{n}=1/\sqrt{3}<p follows by taking the limit nn\to\infty in (3.16).

It follows from the above reasoning that for all xEεx\in\partial E_{\varepsilon} and each (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}) the functions gξ,y𝒢(x)g_{\xi,y}\in\mathcal{G}(x) are 2/32/\sqrt{3}-Lipschitz, since for any w,s[0,r]w,s\in[0,r]

gξ,y(s)gξ,y(w)\displaystyle\|g_{\xi,y}(s)-g_{\xi,y}(w)\| =|sw|2+ε2|fξ,y(s)fξ,y(w)|223|sw|.\displaystyle=\sqrt{|s-w|^{2}+\varepsilon^{2}|f^{\xi,y}(s)-f^{\xi,y}(w)|^{2}}\leq\frac{2}{\sqrt{3}}|s-w|.\qed

Refer to caption

Figure 8. The slopes D±hξ,y(s)D^{\pm}h^{\xi,y}(s) are bounded for all s(0,r)s\in(0,r) from above by the ratio Dmax=a/(ε/2)D_{\mathrm{max}}=a/(\varepsilon/2), which is equal to (ε/2)/b=(ε/2)/ε2(ε/2)2=1/3(\varepsilon/2)/b=(\varepsilon/2)/\sqrt{\varepsilon^{2}-(\varepsilon/2)^{2}}=1/\sqrt{3}. A lower bound is given similarly by D±fξ,y(s)Dmin=1/3D^{\pm}f^{\xi,y}(s)\geq D_{\mathrm{min}}=-1/\sqrt{3}.

3.3. Local Structure of the Complement

In this section we analyse the connectedness of the complement EεcE_{\varepsilon}^{c} near points xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E) and wedges. For these points Lemma 3.8 guarantees the existence of a unique connected component VEεcV\subset E_{\varepsilon}^{c} for which xVx\in\partial V, while Proposition 3.9 makes the stronger statement that in fact EεcBr(x)=VBr(x)E_{\varepsilon}^{c}\cap B_{r}(x)=V\cap B_{r}(x) for some connected VEεcV\subset E_{\varepsilon}^{c} and neighbourhood Br(x)B_{r}(x).

Lemma 3.7 (Approximations of the outward cone).

Let xEεx\in\partial E_{\varepsilon}, let ξ1,ξ2intS1Ξx(Eε)\xi_{1},\xi_{2}\in\textrm{\emph{int}}_{S^{1}}\Xi_{x}(E_{\varepsilon}) and define for each r>0r>0 the truncated cone

Vr(x,ξ1,ξ2):={x+sv:v(ξ1,ξ2)S1, 0<s<r}.V_{r}(x,\xi_{1},\xi_{2}):=\left\{x+sv\,:\,v\in(\xi_{1},\xi_{2})_{S^{1}},\,0<s<r\right\}.

Then there exists r>0r>0 for which Vr(x,ξ1,ξ2)EεcV_{r}(x,\xi_{1},\xi_{2})\subset E_{\varepsilon}^{c}.

Proof.

Assume to the contrary that for each nn\in\mathbb{N} there exists znEεV1/n(x,ξ1,ξ2)z_{n}\in E_{\varepsilon}\cap V_{1/n}(x,\xi_{1},\xi_{2}). Then znxz_{n}\to x as nn\to\infty and we may assume without loss of generality that

znxznxξ[ξ1,ξ2]S1.\frac{z_{n}-x}{\left\lVert z_{n}-x\right\rVert}\to\xi\in[\xi_{1},\xi_{2}]_{S^{1}}.

Since ξ1,ξ2intS1Ξx(Eε)\xi_{1},\xi_{2}\in\textrm{int}_{S^{1}}\Xi_{x}(E_{\varepsilon}), Proposition 2.12 implies ξintS1Ξx(Eε)\xi\in\textrm{int}_{S^{1}}\Xi_{x}(E_{\varepsilon}), which together with Lemma 2.10 (i) gives yx,ξ<0\langle y-x,\xi\rangle<0 for all yΠE(x)y\in\Pi_{E}(x). It follows then from Lemma 2.9 (ii) that there exists NN\in\mathbb{N}, for which znEεcz_{n}\in E_{\varepsilon}^{c} for all nNn\geq N, which contradicts the assumption. ∎

Lemma 3.8 (Unique connected component).

Let E2E\subset\mathbb{R}^{2} and let xEεx\in\partial E_{\varepsilon} either be a wedge (type S1) or xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E). Then there exists a unique connected component VEεcV\subset E_{\varepsilon}^{c}, for which xVx\in\partial V.

Proof.

Since intS1Ξ(x)\textrm{int}_{S^{1}}\Xi(x)\neq\varnothing whenever xx is a wedge or xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E), there exist outward directions ξ1,ξ2intS1Ξx(Eε)\xi_{1},\xi_{2}\in\textrm{int}_{S^{1}}\Xi_{x}(E_{\varepsilon}). Lemma 3.7 then guarantees the existence of some r>0r>0 for which the truncated open cone

(3.17) Vr(x,ξ1,ξ2):={x+sv:v(ξ1,ξ2)S1, 0<s<r}V_{r}(x,\xi_{1},\xi_{2}):=\left\{x+sv\,:\,v\in(\xi_{1},\xi_{2})_{S^{1}},\,0<s<r\right\}

satisfies Vr(x,ξ1,ξ2)EεcV_{r}(x,\xi_{1},\xi_{2})\subset E_{\varepsilon}^{c}. Consequently there exists a connected component VEεcV\subset E_{\varepsilon}^{c} with Vr(x,ξ1,ξ2)VV_{r}(x,\xi_{1},\xi_{2})\subset V and xVx\in\partial V.

For each (ξ,y)𝒫xext(Eε)(\xi,y)\in\mathcal{P}_{x}^{\mathrm{ext}}(E_{\varepsilon}), let fξ,y:[0,ε/2]f^{\xi,y}:[0,\varepsilon/2]\to\mathbb{R} be the continuous function corresponding to the local boundary representation 𝒢(x)\mathcal{G}(x) (see equation 3.8 in Proposition 3.5). To prove uniqueness, assume contrary to the claim that there exists another connected component WEεcW\subset E_{\varepsilon}^{c} with WVW\neq V and xWx\in\partial W. Then for at least one ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) there exist arbitrarily small coordinates s>0s>0 for which

w(s):=x+sξ+tW(s)(xy)Ww(s):=x+s\xi+t_{W}(s)(x-y)\in W

for some tW(s)>fξ,y(s)t_{W}(s)>f^{\xi,y}(s). Since the outward directions ξ1,ξ2\xi_{1},\xi_{2} in the definition of the cone Vr(x,ξ1,ξ2)V_{r}(x,\xi_{1},\xi_{2}) above may be chosen arbitrarily close to the extremal outward directions, it follows that for all sufficiently small ss there also exist coordinates tV(s)>tW(s)t_{V}(s)>t_{W}(s) for which

v(s):=x+sξ+tV(s)(xy)Vr(x,ξ1,ξ2)V.v(s):=x+s\xi+t_{V}(s)(x-y)\in V_{r}(x,\xi_{1},\xi_{2})\subset V.

It follows then from tW(s)>fξ,y(s)t_{W}(s)>f^{\xi,y}(s) and the definition of the local boundary representation that in fact

z(s):=x+sξ+t(xy)Eεcz(s):=x+s\xi+t(x-y)\in E_{\varepsilon}^{c}

for all t[tW(s),tV(s)]t\in[t_{W}(s),t_{V}(s)]. But this contradicts the assumption that VV and WW are both connected and VW=V\cap W=\varnothing. ∎

Proposition 3.9 (Geometry of the complement).

Let E2E\subset\mathbb{R}^{2} and let xEεx\in\partial E_{\varepsilon} either be a wedge or xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E). Then there exists some r>0r>0, for which

(3.18) EεcBr(x)=VBr(x)=0<ρ<rx+A(ρ),E_{\varepsilon}^{c}\cap B_{r}(x)=V\cap B_{r}(x)=\bigcup_{0<\rho<r}x+A(\rho),

where VEεcV\subset E_{\varepsilon}^{c} is connected and for each ρ(0,r)\rho\in(0,r) either A(ρ)=ρ(αρ,βρ)S1A(\rho)=\rho(\alpha_{\rho},\beta_{\rho})_{S^{1}} or A(ρ)=ρ(S1[αρ,βρ]S1)A(\rho)=\rho\big{(}S^{1}\setminus[\alpha_{\rho},\beta_{\rho}]_{S^{1}}\big{)} for αρ,βρS1\alpha_{\rho},\beta_{\rho}\in S^{1} and αρξα\alpha_{\rho}\to\xi_{\alpha}, βρξβ\beta_{\rho}\to\xi_{\beta}, where Ξext(x)={ξα,ξβ}\Xi_{\textrm{ext}}(x)=\{\xi_{\alpha},\xi_{\beta}\}.

Proof.

According to Proposition 3.5 we may assume that xx has a local boundary representation 𝒢(x)\mathcal{G}(x) with radius r>0r>0 for which the functions gξ,y𝒢(x)g_{\xi,y}\in\mathcal{G}(x) are of the form

gξ,y(s)=x+sξ+fξ,y(s)(xy)g_{\xi,y}(s)=x+s\xi+f^{\xi,y}(s)(x-y)

with the functions fξ,y:[0,ε/2]f^{\xi,y}:[0,\varepsilon/2]\to\mathbb{R} continuous. We divide the proof into two parts depending on whether xx is a wedge or xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E).

(i) Assume first that xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E) with ΠE(x)={y}\Pi_{E}(x)=\{y\}. Then 𝒢(x)={gξ1,y,gξ2,y}\mathcal{G}(x)=\left\{g_{\xi_{1},y},g_{\xi_{2},y}\right\} where gξi,y:[0,r]2g_{\xi_{i},y}\,:\,[0,r]\to\mathbb{R}^{2} and

gξi,y(s)=x+sξi+fξi,y(s)(xy)g_{\xi_{i},y}(s)=x+s\xi_{i}+f^{\xi_{i},y}(s)(x-y)

for i{1,2}i\in\{1,2\}. Consider for each s(0,r)s\in(0,r) the distances

𝒟1(s):=gξ1,y(s)x,𝒟2(s):=gξ2,y(s)x.\mathcal{D}_{1}(s):=\left\lVert g_{\xi_{1},y}(s)-x\right\rVert,\quad\mathcal{D}_{2}(s):=\left\lVert g_{\xi_{2},y}(s)-x\right\rVert.

Due to Proposition 2.14 and Lemma 2.13 (i) we can assume rr to be small enough so that both 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are strictly increasing on (0,r)(0,r). One can thus define for each ρ(0,r)\rho\in(0,r) the points αρ,βρS1\alpha_{\rho},\beta_{\rho}\in S^{1} by

αρ:=ρ1(gξ1,y(𝒟11(ρ))x),βρ:=ρ1(gξ2,y(𝒟21(ρ))x).\displaystyle\alpha_{\rho}:=\rho^{-1}\left(g_{\xi_{1},y}\left(\mathcal{D}_{1}^{-1}(\rho)\right)-x\right),\qquad\beta_{\rho}:=\rho^{-1}\left(g_{\xi_{2},y}\left(\mathcal{D}_{2}^{-1}(\rho)\right)-x\right).

According to Lemma 3.8 there exists a unique connected component VV of EεcE_{\varepsilon}^{c} for which xVx\in\partial V, and due to Proposition 3.1 we can assume rr to be sufficiently small so that

Br(x)EBε(EBε/2(y))¯.B_{r}(x)\cap E\subset\overline{B_{\varepsilon}\left(E\cap B_{\varepsilon/2}(y)\right)}.

This implies that for each ρ(0,r)\rho\in(0,r) the geodesic curve segment

A(ρ):=ρ(αρ,(xy)/ε)S1{ρ(xy)/ε}ρ((xy)/ε,βρ)S1ρS1A(\rho):=\rho(\alpha_{\rho},(x-y)/\varepsilon)_{S^{1}}\cup\{\rho(x-y)/\varepsilon\}\cup\rho((x-y)/\varepsilon,\beta_{\rho})_{S^{1}}\subset\rho S^{1}

satisfies

x+A(ρ)VBr(x)andx+ρS1A(ρ)Bε(EBε/2(y))¯Eε.x+A(\rho)\subset V\cap B_{r}(x)\quad\mathrm{and}\quad x+\rho S^{1}\setminus A(\rho)\subset\overline{B_{\varepsilon}\left(E\cap B_{\varepsilon/2}(y)\right)}\subset E_{\varepsilon}.

Hence

EεcBr(x)=VBr(x)=0<ρ<rx+A(ρ).E_{\varepsilon}^{c}\cap B_{r}(x)=V\cap B_{r}(x)=\bigcup_{0<\rho<r}x+A(\rho).

(ii) Let then xx be a wedge. As above, one can assume that the distances

𝒟1(s):=gξ1,y1(s)x,𝒟2(s):=gξ2,y2(s)x\mathcal{D}_{1}(s):=\left\lVert g_{\xi_{1},y_{1}}(s)-x\right\rVert,\quad\mathcal{D}_{2}(s):=\left\lVert g_{\xi_{2},y_{2}}(s)-x\right\rVert

are strictly increasing in (0,r)(0,r). Furthermore, since ξ1,ξ2>1\langle\xi_{1},\xi_{2}\rangle>-1, one may define the average ξav:=(ξ1+ξ2)/ξ1+ξ2\xi_{\mathrm{av}}:=(\xi_{1}+\xi_{2})/\left\lVert\xi_{1}+\xi_{2}\right\rVert. Due to Lemma 2.13 (i) and that fact that ξavΞxext(Eε)\xi_{\mathrm{av}}\notin\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}), we may assume rr to be small enough so that the boundary segments represented by the functions gξi,yig_{\xi_{i},y_{i}} are separated by the line segment {x+ρξav:ρ(0,r)}\{x+\rho\xi_{\mathrm{av}}\,:\,\rho\in(0,r)\} in the neighbourhood Br(x)B_{r}(x).

For each ρ(0,r)\rho\in(0,r), we once again define the points αρ,βρS1\alpha_{\rho},\beta_{\rho}\in S^{1} by

αρ:=ρ1(gξ1,y(𝒟11(ρ))x)andβρ:=ρ1(gξ2,y(𝒟21(ρ))x).\displaystyle\alpha_{\rho}:=\rho^{-1}\left(g_{\xi_{1},y}\left(\mathcal{D}_{1}^{-1}(\rho)\right)-x\right)\qquad\textrm{and}\qquad\beta_{\rho}:=\rho^{-1}\left(g_{\xi_{2},y}\left(\mathcal{D}_{2}^{-1}(\rho)\right)-x\right).

Analogously to part (i), Proposition 3.1 and Lemma 3.8 guarantee that for all ρ(0,r)\rho\in(0,r) the geodesic curve segment

A^(ρ):=ρ(αρ,ξav)S1{ρξav}ρ(ξav,βρ)S1ρS1\widehat{A}(\rho):=\rho(\alpha_{\rho},\xi_{\mathrm{av}})_{S^{1}}\cup\{\rho\xi_{\mathrm{av}}\}\cup\rho(\xi_{\mathrm{av}},\beta_{\rho})_{S^{1}}\subset\rho S^{1}

satisfies

x+A^(ρ)VBr(x)andx+ρS1A^(ρ)Bε(EBε/2({y1,y2}))¯Eϵ,x+\widehat{A}(\rho)\subset V\cap B_{r}(x)\quad\mathrm{and}\quad x+\rho S^{1}\setminus\widehat{A}(\rho)\subset\overline{B_{\varepsilon}\left(E\cap B_{\varepsilon/2}(\{y_{1},y_{2}\})\right)}\subset E_{\epsilon},

where VV is the unique connected component of the complement EεcE_{\varepsilon}^{c} for which Br(x)Eεc=Br(x)VB_{r}(x)\cap E_{\varepsilon}^{c}=B_{r}(x)\cap V. Hence

EεcBr(x)=VBr(x)=0<ρ<rx+A^(ρ).E_{\varepsilon}^{c}\cap B_{r}(x)=V\cap B_{r}(x)=\bigcup_{0<\rho<r}x+\widehat{A}(\rho).\qed

4. Classification of Boundary Points

In this section we present a classification of the boundary points xEεx\in\partial E_{\varepsilon} based on their local geometric and topological properties. Using the results obtained in Sections 2 and 3 above, we prove our first main result, Theorem 1, which states that the classification given in Definition 4.1 defines a partition of the boundary Eε\partial E_{\varepsilon} into disjoint subsets.

The geometric aspect of the classification scheme relies on the orientation of the extremal contributors yΠEext(x)y\in\Pi_{E}^{\mathrm{ext}}(x) at each boundary point xEεx\in\partial E_{\varepsilon}. In the planar case, there are essentially three different ways this orientation can be realised, depicted schematically in Figure 9 below. The defining property y1x=(y2x)y_{1}-x=-(y_{2}-x) for the extremal contributors y1,y2ΠEext(x)y_{1},y_{2}\in\Pi_{E}^{\mathrm{ext}}(x) in case (c) can be equivalently expressed by (y1x)/ε,(y2x)/ε=1\big{\langle}(y_{1}-x)/\varepsilon,(y_{2}-x)/\varepsilon\big{\rangle}=-1, and we will make use of both formulations in what follows.

Refer to caption
(a) At each xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E) the extremal outward directions satisfy ξ1=ξ2\xi_{1}=-\xi_{2}, while the set Ξx(Eε)\Xi_{x}(E_{\varepsilon}) spans a half-circle.
Refer to caption
(b) For a wedge xx there are two extremal contributors y1,y2y_{1},y_{2} and two extremal outward directions ξ1,ξ2\xi_{1},\xi_{2}, forming an angle θ=(ξ1,ξ2)\theta=\sphericalangle(\xi_{1},\xi_{2}).
Refer to caption
(c) For ΠEext(x)={y1,y2}\Pi_{E}^{\mathrm{ext}}(x)=\{y_{1},y_{2}\} with y1x=(y2x)y_{1}-x=-(y_{2}-x), the set of extremal outward directions satisfies Ξxext(Eε){ξ1,ξ2}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})\subset\{\xi_{1},\xi_{2}\}.
Figure 9. The local geometry at each boundary point xEεx\in\partial E_{\varepsilon} reflects the three basic scenarios (a)–(c) regarding the number and positions of contributors yΠE(x)y\in\Pi_{E}(x). In our classification of boundary points (see Definitions 2.1, 2.2 and 4.1), case (a) corresponds to smooth points and singularities of types S4 and S5, case (b) to type S1, and case (c) to types S2, S3 and S6–S8. See also Figure 1 and Proposition 2.12 regarding the structure of the set of outward directions Ξx(Eε)\Xi_{x}(E_{\varepsilon}).

4.1. Types of Singularities

The classification of singularities is given in Definition 4.1 below. Schematic illustrations of the different types of singularities are given in Figure 1 in the Introduction. Recall that Ur(x,v)U_{r}(x,v) denotes an open xx-centered half-ball of radius rr oriented in the direction of vS1v\in S^{1} (see (3.2)). We denote by S(Eε)S(E_{\varepsilon}) the set of singularities on the boundary Eε\partial E_{\varepsilon}.

Definition 4.1 (Types of singularities).

Let E2E\subset\mathbb{R}^{2} be closed, let xS(Eε)x\in S(E_{\varepsilon}) and let Ξxext(Eε)={ξ1,ξ2}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi_{1},\xi_{2}\} be the set of extremal outward directions, where we allow for the possibility ξ1=ξ2\xi_{1}=\xi_{2}. We define the following eight types of singularities.

  1. S1:

    xx is a wedge, if ξ1{ξ2,ξ2}\xi_{1}\notin\{\xi_{2},-\xi_{2}\}, i.e. the angle θ\theta between the vectors ξ1,ξ2\xi_{1},\xi_{2} satisfies 0<θ<π0<\theta<\pi;

  2. S2:

    xx is a (one-sided) sharp singularity, if ξ1=ξ2\xi_{1}=\xi_{2}, and there exists some δ>0\delta>0 for which the intersection Bδ(x)EεcB_{\delta}(x)\cap E_{\varepsilon}^{c} is a connected set;

  3. S3:

    xx is a sharp-sharp singularity, if ξ1=ξ2\xi_{1}=-\xi_{2} and for each i{1,2}i\in\{1,2\} there exists some δi>0\delta_{i}>0 for which the intersection Uδi(x,ξi)EεcU_{\delta_{i}}(x,\xi_{i})\cap E_{\varepsilon}^{c} is a connected set;

  4. S4:

    xx is a (one-sided) shallow singularity if xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E) and

    1. (i)

      Uδ1(x,ξ1)EεUnpε(E)U_{\delta_{1}}(x,\xi_{1})\cap\partial E_{\varepsilon}\subset\mathrm{Unp_{\varepsilon}}(E) for some δ1>0\delta_{1}>0, and

    2. (ii)

      Uδ2(x,ξ2)EεUnpε(E)U_{\delta_{2}}(x,\xi_{2})\cap\partial E_{\varepsilon}\not\subset\mathrm{Unp_{\varepsilon}}(E) for all δ2>0\delta_{2}>0.

  5. S5:

    xx is a shallow-shallow singularity if xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E) and Uδ(x,ξi)EεUnpε(E)U_{\delta}(x,\xi_{i})\cap\partial E_{\varepsilon}\not\subset\mathrm{Unp_{\varepsilon}}(E) for all δ>0\delta>0 and i{1,2}i\in\{1,2\}.

  6. S6:

    xx is a (one-sided) chain singularity, if ξ1=ξ2\xi_{1}=\xi_{2} and there exists a sequence of singularities (xn)n=1S(Eε)(x_{n})_{n=1}^{\infty}\subset S(E_{\varepsilon}), for which xnxx_{n}\to x and

    yn(1)xnε,yn(2)xnε1,\left\langle\frac{y_{n}^{(1)}-x_{n}}{\varepsilon},\frac{y_{n}^{(2)}-x_{n}}{\varepsilon}\right\rangle\rightarrow-1,

    where {yn(1),yn(2)}=ΠEext(xn)\big{\{}y_{n}^{(1)},y_{n}^{(2)}\big{\}}=\Pi_{E}^{\mathrm{ext}}(x_{n}) is the set of extremal contributors at each xnx_{n};

  7. S7:

    xx is a chain-chain singularity, if ξ1=ξ2\xi_{1}=-\xi_{2} and for each i{1,2}i\in\{1,2\} there exists some δi>0\delta_{i}>0 and a sequence (xi,n)n=1Uδi(x,ξi)S(Eε)(x_{i,n})_{n=1}^{\infty}\subset U_{\delta_{i}}(x,\xi_{i})\cap S(E_{\varepsilon}), for which xi,nxx_{i,n}\to x and

    yi,n(1)xi,nε,yi,n(2)xi,nε1,\left\langle\frac{y_{i,n}^{(1)}-x_{i,n}}{\varepsilon},\frac{y_{i,n}^{(2)}-x_{i,n}}{\varepsilon}\right\rangle\rightarrow-1,

    where {yi,n(1),yi,n(2)}=ΠEext(xi,n)\big{\{}y_{i,n}^{(1)},y_{i,n}^{(2)}\big{\}}=\Pi_{E}^{\mathrm{ext}}(x_{i,n}) is the set of extremal contributors at each xi,nx_{i,n};

  8. S8:

    xx is a sharp-chain singularity, if ξ1=ξ2\xi_{1}=-\xi_{2} and

    1. (i)

      there exists a δ1>0\delta_{1}>0 for which the intersection Uδ1(x,ξ1)EεcU_{\delta_{1}}(x,\xi_{1})\cap E_{\varepsilon}^{c} is a connected set, and

    2. (ii)

      there exists some δ2>0\delta_{2}>0 and a sequence (xn)n=1Uδ2(x,ξ2)S(Eε)(x_{n})_{n=1}^{\infty}\subset U_{\delta_{2}}(x,\xi_{2})\cap S(E_{\varepsilon}), for which xnxx_{n}\to x and

      yn(1)xnε,yn(2)xnε1,\left\langle\frac{y_{n}^{(1)}-x_{n}}{\varepsilon},\frac{y_{n}^{(2)}-x_{n}}{\varepsilon}\right\rangle\rightarrow-1,

      where {yn(1),yn(2)}=ΠEext(xn)\big{\{}y_{n}^{(1)},y_{n}^{(2)}\big{\}}=\Pi_{E}^{\mathrm{ext}}(x_{n}) is the set of extremal contributors at each xnx_{n}.

Note that S8 may be interpreted both as a sharp singularity and as a chain singularity. Theorem 3 below states that the set 𝒞(Eε):={xEε:x is of type S6–S8}\mathcal{C}(\partial E_{\varepsilon}):=\{x\in\partial E_{\varepsilon}\,:\,x\textrm{ is of type S6--S8}\} is closed, while on the other hand all the singularities of type S8 share an important property with those of type S1–S5: they all lie on the boundary V\partial V of some connected component VV of the complement EεcE_{\varepsilon}^{c}. We show in Corollary 4.3 that this is exactly the property that is lacking from singularities of type S6 and S7 (see also Remark 4.4 below).

Motivated by these considerations we define a boundary point to be

  • (i)

    a sharp singularity, if it is of type S2, S3 or S8,

  • (ii)

    a chain singularity, if it is of type S6, S7 or S8, and

  • (iii)

    an inaccessible singularity, if it is of type S6 or S7.

The typology presented above is neither strictly topological nor strictly geometric. If one wanted to accomplish a strictly topological classification for neighbourhoods EεBδ(x)\partial E_{\varepsilon}\cap B_{\delta}(x) for some δ:=δ(x)>0\delta:=\delta(x)>0, types S6–S8 would necessitate an infinite tree-like classification scheme, in order to account for the potentially accumulating chain and shallow singularities in arbitrarily small neighbourhoods Br(x)B_{r}(x) with 0<r<δ0<r<\delta (see Section 5.2 and Theorem 3).

4.2. Classification of Singularities

Proposition 4.2 below provides a characterisation of the topological and geometric structure of the complement EεcE_{\varepsilon}^{c} near those singularities xS(Eε)x\in S(E_{\varepsilon}), whose extremal contributors y1,y2y_{1},y_{2} satisfy y1x=(y2x)y_{1}-x=-(y_{2}-x). Geometrically these correspond to case (c) in Figure 9.

Proposition 4.2 (Difference between sharp-type and chain-type geometry).

Let E2E\subset\mathbb{R}^{2}, xEεx\in\partial E_{\varepsilon} and ΠEext(x)={y1,y2}\Pi_{E}^{\mathrm{ext}}(x)=\{y_{1},y_{2}\} with y1x=(y2x)y_{1}-x=-(y_{2}-x). Furthermore, let 𝒢(x)\mathcal{G}(x) be a local boundary representation with radius r>0r>0 at xx, let ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) and let gξ,y1,gξ,y2𝒢(x)g_{\xi,y_{1}},g_{\xi,y_{2}}\in\mathcal{G}(x) be as in (3.8). Then exactly one of the cases (i) and (ii) below holds true:

  • (i)

    (sharp-type) There exists some r>0r>0, for which gξ,y1(s)gξ,y2(s)g_{\xi,y_{1}}(s)\neq g_{\xi,y_{2}}(s) for all s(0,r)s\in(0,r), and

    (4.1) EεcUr(x,ξ)=VξUr(x,ξ)=0<s<rx+s(α(s),β(s))S1,E_{\varepsilon}^{c}\cap U_{r}(x,\xi)=V_{\xi}\cap U_{r}(x,\xi)=\bigcup_{0<s<r}x+s\big{(}\alpha(s),\beta(s)\big{)}_{S^{1}},

    where VξV_{\xi} is the unique connected component of EεcE_{\varepsilon}^{c} intersecting Ur(x,ξ)U_{r}(x,\xi), α(s),β(s)S1\alpha(s),\beta(s)\in S^{1} for all s(0,r)s\in(0,r) and α(s),β(s)ξ\alpha(s),\beta(s)\to\xi as s0s\to 0.

  • (ii)

    (chain-type) There exists a sequence (sn)n=1+(s_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} with the following properties:

    • (a)

      sn0s_{n}\to 0 and gξ,y1(sn)=gξ,y2(sn)g_{\xi,y_{1}}(s_{n})=g_{\xi,y_{2}}(s_{n}) for all nn\in\mathbb{N}. We denote this common value by xnx_{n}.

    • (b)

      There exists some r>0r>0 and a sequence (Vn)n=1Ur(x,ξ)(V_{n})_{n=1}^{\infty}\subset U_{r}(x,\xi) of disjoint connected components of EεcE_{\varepsilon}^{c} with distH(x,Vn)0\mathrm{dist}_{H}(x,V_{n})\to 0 as nn\to\infty and xnVnx_{n}\in\partial V_{n} for all nn\in\mathbb{N}.

    • (c)

      xnS(Eε)x_{n}\in S(E_{\varepsilon}) for each nn\in\mathbb{N}, with

      (4.2) limnyn(1)xnε,yn(2)xnε=1,\lim_{n\to\infty}\left\langle\frac{y_{n}^{(1)}-x_{n}}{\varepsilon},\frac{y_{n}^{(2)}-x_{n}}{\varepsilon}\right\rangle=-1,

      where ΠEext(xn)={yn(1),yn(2)}\Pi_{E}^{\mathrm{ext}}(x_{n})=\big{\{}y_{n}^{(1)},y_{n}^{(2)}\big{\}} for all nn\in\mathbb{N}.

Proof.

Clearly, either there exists some r>0r>0, for which gξ,y1(s)gξ,y2(s)g_{\xi,y_{1}}(s)\neq g_{\xi,y_{2}}(s) for all s(0,r)s\in(0,r), or else there exists a sequence (qn)n=1+(q_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} with qn0q_{n}\to 0, for which gξ,y1(qn)=gξ,y2(qn)g_{\xi,y_{1}}(q_{n})=g_{\xi,y_{2}}(q_{n}) for all nn\in\mathbb{N}, and these cases are mutually exclusive. The proof amounts to showing that in the former case representation (4.1) is valid for some connected component VξEεcV_{\xi}\subset E_{\varepsilon}^{c} and arc-segments (α(s),β(s))S1(\alpha(s),\beta(s))_{S^{1}}, and in the latter, to identifying the prescribed sequences (sn)n=1+(s_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} and (Vn)n=1Ur(x,ξ)(V_{n})_{n=1}^{\infty}\subset U_{r}(x,\xi) as well as confirming the limit (4.2) and that xnVnx_{n}\in\partial V_{n} for all nn\in\mathbb{N}.

Consider for i{1,2}i\in\{1,2\} the continuous functions fξ,yi:[0,ε/2]f^{\xi,y_{i}}:[0,\varepsilon/2]\to\mathbb{R} for which

gξ,y1(s)=x+sξ+fξ,y1(s)(xy1)x+sξ+fξ,y2(s)(xy2)=gξ,y2(s)g_{\xi,y_{1}}(s)=x+s\xi+f^{\xi,y_{1}}(s)(x-y_{1})\neq x+s\xi+f^{\xi,y_{2}}(s)(x-y_{2})=g_{\xi,y_{2}}(s)

(see Proposition 3.5). The assumption xy1=(xy2)x-y_{1}=-(x-y_{2}) implies that the vector representing the difference at s(0,r)s\in(0,r) between the graphs gξ,y1([0,r])g_{\xi,y_{1}}([0,r]) and gξ,y2([0,r])g_{\xi,y_{2}}([0,r]), is given by

(4.3) gξ,y2(s)gξ,y1(s)=(fξ,y1(s)+fξ,y2(s))(xy1).g_{\xi,y_{2}}(s)-g_{\xi,y_{1}}(s)=-\left(f^{\xi,y_{1}}(s)+f^{\xi,y_{2}}(s)\right)(x-y_{1}).

(i) We start by assuming that there exists some r>0r>0, for which gξ,y1(s)gξ,y2(s)g_{\xi,y_{1}}(s)\neq g_{\xi,y_{2}}(s) for all s(0,r)s\in(0,r) which implies αξ(s):=fξ,y1(s)+fξ,y2(s)0\alpha_{\xi}(s):=f^{\xi,y_{1}}(s)+f^{\xi,y_{2}}(s)\neq 0 for all s(0,r)s\in(0,r). Due to continuity, this implies either

(1)αξ(s)>0for alls(0,r),or(2)αξ(s)<0for alls(0,r).\textrm{(1)}\quad\alpha_{\xi}(s)>0\,\,\textrm{for all}\,\,s\in(0,r),\qquad\textrm{or}\qquad\textrm{(2)}\quad\alpha_{\xi}(s)<0\,\,\textrm{for all}\,\,s\in(0,r).

Note that (1) would contradict the assumption ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}), so that (2) necessarily holds true. Hence the average hξ:[0,r]2h_{\xi}:[0,r]\to\mathbb{R}^{2}, given by

(4.4) hξ(s):=gξ,y1(s)+gξ,y2(s)2=x+sξ+(fξ,y1(s)αξ(s)2)(xy1),h_{\xi}(s):=\frac{g_{\xi,y_{1}}(s)+g_{\xi,y_{2}}(s)}{2}=x+s\xi+\left(f^{\xi,y_{1}}(s)-\frac{\alpha_{\xi}(s)}{2}\right)(x-y_{1}),

satisfies hξ(s)Eεch_{\xi}(s)\in E_{\varepsilon}^{c} for all s(0,r)s\in(0,r). Note that although we have defined the function hξh_{\xi} in (4.4) in terms of the contributor y1y_{1}, we could have equally well chosen y2y_{2} due to symmetry. Combining the facts that hξh_{\xi} is continuous, hξ((0,r))Eεch_{\xi}((0,r))\subset E_{\varepsilon}^{c} and x=hξ(0)x=h_{\xi}(0), we may deduce that there exists a connected component VξEεcV_{\xi}\subset E_{\varepsilon}^{c} for which hξ((0,r))Vξh_{\xi}((0,r))\subset V_{\xi} and xVξx\in\partial V_{\xi}. Emulating the reasoning presented in the proof of Lemma 3.8 allows one to confirm that VξV_{\xi} is the only connected component of EεcE_{\varepsilon}^{c} that intersects Ur(x,ξ)U_{r}(x,\xi).

To obtain representation (4.1), consider for each s(0,r)s\in(0,r) the unit vectors h(s),α(s),β(s)S1h(s),\alpha(s),\beta(s)\in S^{1}, given by

(4.5) h(s):=hξ(s)xhξ(s)x,α(s):=gξ,y1(s)xgξ,y1(s)x,β(s):=gξ,y2(s)xgξ,y2(s)x.h(s):=\frac{h_{\xi}(s)-x}{\left\lVert h_{\xi}(s)-x\right\rVert},\quad\alpha(s):=\frac{g_{\xi,y_{1}}(s)-x}{\left\lVert g_{\xi,y_{1}}(s)-x\right\rVert},\quad\beta(s):=\frac{g_{\xi,y_{2}}(s)-x}{\left\lVert g_{\xi,y_{2}}(s)-x\right\rVert}.

Due to Proposition 2.14 and Lemma 2.13 (i) we know that the boundary Eε\partial E_{\varepsilon} aligns itself with the extremal outward directions ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) near each boundary point xEεx\in\partial E_{\varepsilon}. We can hence assume rr to be small enough such that the distances

𝒟h(s):=hξ(s)x,𝒟α(s):=gξ,y1(s)x,𝒟β(s):=gξ,y2(s)x\mathcal{D}_{h}(s):=\left\lVert h_{\xi}(s)-x\right\rVert,\quad\mathcal{D}_{\alpha}(s):=\left\lVert g_{\xi,y_{1}}(s)-x\right\rVert,\quad\mathcal{D}_{\beta}(s):=\left\lVert g_{\xi,y_{2}}(s)-x\right\rVert

appearing in the divisors in (4.5) are all strictly increasing in ss on the interval (0,r)(0,r). By definition we also have max{𝒟h1(s),𝒟α1(s),𝒟β1(s)}s\mathrm{max}\,\big{\{}\mathcal{D}_{h}^{-1}(s),\mathcal{D}_{\alpha}^{-1}(s),\mathcal{D}_{\beta}^{-1}(s)\big{\}}\leq s for all s(0,r)s\in(0,r).

Hence, for each s(0,r)s\in(0,r)

hξ(𝒟h1(s))=x+sh(𝒟h1(s))x+s(α(𝒟α1(s)),β(𝒟β1(s)))S1VξUr(x,ξ).h_{\xi}\big{(}\mathcal{D}_{h}^{-1}(s)\big{)}=x+sh\big{(}\mathcal{D}_{h}^{-1}(s)\big{)}\in x+s\big{(}\alpha(\mathcal{D}_{\alpha}^{-1}(s)),\beta(\mathcal{D}_{\beta}^{-1}(s))\big{)}_{S^{1}}\subset V_{\xi}\cap U_{r}(x,\xi).

In addition, we can assume Proposition 3.1 to apply at xx with the choices r:=rr:=r and δ:=ε/2\delta:=\varepsilon/2. From this it follows for Cs:=s(S1(α(s),β(s))S1)Ur(x,ξ)C_{s}:=s\left(S^{1}\setminus\big{(}\alpha(s),\beta(s)\big{)}_{S^{1}}\right)\cap U_{r}(x,\xi) that

x+CsBε(EBε/2(ΠEext(x)))¯Eεx+C_{s}\subset\overline{B_{\varepsilon}\left(E\cap B_{\varepsilon/2}(\Pi_{E}^{\mathrm{ext}}(x))\right)}\subset E_{\varepsilon}

for all s(0,r)s\in(0,r). Hence

EεcUr(x,ξ)=VξUr(x,ξ)=0<s<rx+s(α(s),β(s))S1.E_{\varepsilon}^{c}\cap U_{r}(x,\xi)=V_{\xi}\cap U_{r}(x,\xi)=\bigcup_{0<s<r}x+s\big{(}\alpha(s),\beta(s)\big{)}_{S^{1}}.

(ii) Assume then that there exists a sequence (qn)n=1+(q_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} for which gξ,y1(qn)=gξ,y2(qn)g_{\xi,y_{1}}(q_{n})=g_{\xi,y_{2}}(q_{n}) for all nn\in\mathbb{N} and qn0q_{n}\to 0. This situation corresponds to the chain-type geometry characteristic of chain singularities (types S6–S8; see Definition 4.1 and Figure 1). Note that since xEεx\in\partial E_{\varepsilon} and ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}), there exists for all s>0s>0 some 0<λ<s0<\lambda<s, for which αξ(λ)=fξ,y1(λ)+fξ,y2(λ)<0\alpha_{\xi}(\lambda)=f^{\xi,y_{1}}(\lambda)+f^{\xi,y_{2}}(\lambda)<0. One can thus define two new sequences (sn)n=1+(s_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} and (pn)n=1+(p_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} inductively as follows. First, choose some λ1(0,q1)\lambda_{1}\in(0,q_{1}) with fξ,y1(λ1)+fξ,y2(λ1)<0f^{\xi,y_{1}}(\lambda_{1})+f^{\xi,y_{2}}(\lambda_{1})<0 and define

s1\displaystyle s_{1} :=sup{s:s>λ1andfξ,y1(λ)+fξ,y2(λ)<0for allλ[λ1,s]},\displaystyle:=\sup\left\{s\,:\,s>\lambda_{1}\,\,\textrm{and}\,\,f^{\xi,y_{1}}(\lambda)+f^{\xi,y_{2}}(\lambda)<0\,\,\textrm{for all}\,\,\lambda\in[\lambda_{1},s]\right\},
p1\displaystyle p_{1} :=inf{s:s<λ1andfξ,y1(λ)+fξ,y2(λ)<0for allλ[s,λ1]}.\displaystyle:=\inf\left\{s\,:\,s<\lambda_{1}\,\,\textrm{and}\,\,f^{\xi,y_{1}}(\lambda)+f^{\xi,y_{2}}(\lambda)<0\,\,\textrm{for all}\,\,\lambda\in[s,\lambda_{1}]\right\}.

For the induction step, assume we have already chosen the points s1,,sn1s_{1},\ldots,s_{n-1} and p1,,pn1p_{1},\ldots,p_{n-1} for some nn\in\mathbb{N}. One can then choose some λn(0,min{pn1,qn})\lambda_{n}\in(0,\textrm{min}\{p_{n-1},q_{n}\}) with fξ,y1(λn)+fξ,y2(λn)<0f^{\xi,y_{1}}(\lambda_{n})+f^{\xi,y_{2}}(\lambda_{n})<0, and define

sn\displaystyle s_{n} :=sup{s:s>λnandfξ,y1(λ)+fξ,y2(λ)<0for allλ[λn,s]},\displaystyle:=\sup\left\{s\,:\,s>\lambda_{n}\,\,\textrm{and}\,\,f^{\xi,y_{1}}(\lambda)+f^{\xi,y_{2}}(\lambda)<0\,\,\textrm{for all}\,\,\lambda\in[\lambda_{n},s]\right\},
pn\displaystyle p_{n} :=inf{s:s<λnandfξ,y1(λ)+fξ,y2(λ)<0for allλ[s,λn]}.\displaystyle:=\inf\left\{s\,:\,s<\lambda_{n}\,\,\textrm{and}\,\,f^{\xi,y_{1}}(\lambda)+f^{\xi,y_{2}}(\lambda)<0\,\,\textrm{for all}\,\,\lambda\in[s,\lambda_{n}]\right\}.

Then pn<snpn1<sn1p_{n}<s_{n}\leq p_{n-1}<s_{n-1} for all nn\in\mathbb{N} and fξ,y1(s)+fξ,y2(s)=0f^{\xi,y_{1}}(s)+f^{\xi,y_{2}}(s)=0 for all s(sn)n=1(pn)n=1s\in(s_{n})_{n=1}^{\infty}\cup(p_{n})_{n=1}^{\infty}. Also, by definition, fξ,y1(s)+fξ,y2(s)<0f^{\xi,y_{1}}(s)+f^{\xi,y_{2}}(s)<0 for all s(pn,sn)s\in(p_{n},s_{n}) and nn\in\mathbb{N}. This implies that for each nn\in\mathbb{N} the open set

Vn:={τgξ,y1(s)+(1τ)gξ,y2(s):τ(0,1),s(pn,sn)}V_{n}:=\left\{\tau g_{\xi,y_{1}}(s)+(1-\tau)g_{\xi,y_{2}}(s)\,:\,\tau\in(0,1),\,s\in(p_{n},s_{n})\right\}

is connected and satisfies xn:=gξ,y1(sn)=gξ,y2(sn)Vnx_{n}:=g_{\xi,y_{1}}(s_{n})=g_{\xi,y_{2}}(s_{n})\in\partial V_{n}. In addition VnVm=V_{n}\cap V_{m}=\varnothing whenever nmn\neq m, and for every r>0r>0 there exists some NN\in\mathbb{N}, for which 0<pn<sn<r0<p_{n}<s_{n}<r for all nNn\geq N. It thus follows from Propositions 2.14 and Lemma 2.13 (i) that distH(x,Vn)0\mathrm{dist}_{H}(x,V_{n})\to 0 as nn\to\infty.

Since xn:=gξ,y1(sn)=gξ,y2(sn)x_{n}:=g_{\xi,y_{1}}(s_{n})=g_{\xi,y_{2}}(s_{n}) for each nn\in\mathbb{N}, there exist for i{1,2}i\in\{1,2\} extremal contributors yn(i)ΠEext(xn)y_{n}^{(i)}\in\Pi_{E}^{\mathrm{ext}}(x_{n}), for which yn(i)Bε/2(yi)y_{n}^{(i)}\in B_{\varepsilon/2}(y_{i}). This, together with Lemma 2.13 (ii)(a), implies yn(i)yiy_{n}^{(i)}\to y_{i} for i{1,2}i\in\{1,2\} as nn\to\infty, and consequently

limnyn(1)xnε,yn(2)xnε=1.\lim_{n\to\infty}\left\langle\frac{y_{n}^{(1)}-x_{n}}{\varepsilon},\frac{y_{n}^{(2)}-x_{n}}{\varepsilon}\right\rangle=-1.\qed

As a consequence of Proposition 4.2 we obtain the following characterisation for inaccessible boundary points xx, which are defined by the property that xVx\notin\partial V for all connected components VV of the complement Eεc:=2EεE_{\varepsilon}^{c}:=\mathbb{R}^{2}\setminus E_{\varepsilon}.

Corollary 4.3 (Inaccessible singularities).

Let E2E\subset\mathbb{R}^{2} and xEεx\in\partial E_{\varepsilon}. Then xVx\notin\partial V for all connected components VV of the complement EεcE_{\varepsilon}^{c} if and only if xx is a one-sided chain singularity (S6) or a chain-chain singularity (S7).

Proof.

Assume first that xVx\notin\partial V for all connected components VV of the complement EεcE_{\varepsilon}^{c}. Proposition 3.9 then implies that xx is not a wedge and xUnpε(E)x\notin\mathrm{Unp_{\varepsilon}}(E), so that the extremal contributors y1,y2ΠEext(x)y_{1},y_{2}\in\Pi_{E}^{\mathrm{ext}}(x) satisfy y1x=(y2x)y_{1}-x=-(y_{2}-x).

In case xx has only one extremal outward direction ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}), Proposition 3.1 implies the existence of some r<ε/2r<\varepsilon/2 for which EεcBr(x)Ur(x,ξ)E_{\varepsilon}^{c}\cap B_{r}(x)\subset U_{r}(x,\xi). By assumption there cannot exist any connected component VξV_{\xi} described in case (i) of Proposition 4.2, which implies that case (ii) holds. Hence xx is a (one-sided) chain singularity.

If, on the other hand, there exist extremal outward directions ξ1,ξ2\xi_{1},\xi_{2} with ξ1=ξ2\xi_{1}=-\xi_{2}, Proposition 4.2 again rules out case (i) for each one of them, and consequently xx fulfils the definition of a chain-chain singularity.

Assume then that xx is either a one-sided chain singularity (S6) or a chain-chain singularity (S7). In the former case Ξxext(Eε)={ξ}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi\} for some ξS1\xi\in S^{1}, and Proposition 3.1 again implies the existence of some r<ε/2r<\varepsilon/2 for which EεcBr(x)Ur(x,ξ)E_{\varepsilon}^{c}\cap B_{r}(x)\subset U_{r}(x,\xi). We aim to deduce a contradiction by assuming there exists a connected VUr(x,ξ)EεcV\subset U_{r}(x,\xi)\cap E_{\varepsilon}^{c} for which xVx\in\partial V. Note that every zVz\in V has a representation

(4.6) z=x+sξ+t(s)(xy1)z=x+s\xi+t(s)(x-y_{1})

for some s=s(z)>0s=s(z)>0 and t(s)t(s)\in\mathbb{R}. Now choose some s0>0s_{0}>0 and t(s0)t(s_{0}) so that z0=x+s0ξ+t(s0)(xy1)Vz_{0}=x+s_{0}\xi+t(s_{0})(x-y_{1})\in V. Given that VV is connected and xVx\in\partial V, it follows from z0Vz_{0}\in V that there exists a path γ:[0,1]V\gamma:[0,1]\to V for which γ(0)=x\gamma(0)=x and γ(1)=z0\gamma(1)=z_{0}. Following (4.6), we may write

γ(u)=x+s(u)ξ+t(s(u))(xy1),\gamma(u)=x+s(u)\xi+t(s(u))(x-y_{1}),

where the coordinate s(u)s(u) depends continuously on u[0,1]u\in[0,1]. Since γ(u)Eε\gamma(u)\notin E_{\varepsilon} for all u[0,1]u\in[0,1], Proposition 3.5 implies that the coordinate t(s(u))t(s(u)) satisfies t(s(u))(fξ,y1(s(u)),fξ,y2(s(u)))t(s(u))\in\big{(}f^{\xi,y_{1}}(s(u)),-f^{\xi,y_{2}}(s(u))\big{)} for all u[0,1]u\in[0,1], where the functions fξ,yif^{\xi,y_{i}} are as in Proposition 3.5. However, Proposition 4.2 implies the existence of some 0<q<s00<q<s_{0} for which fξ,y1(q)=fξ,y2(q)f^{\xi,y_{1}}(q)=-f^{\xi,y_{2}}(q), and since s()s(\cdot) is continuous as a function of uu, there exists some uqu_{q} for which s(uq)=qs(u_{q})=q. For this coordinate we thus obtain the contradiction t(s(uq))=t(q)(fξ,y1(q),fξ,y2(q))=t(s(u_{q}))=t(q)\in\big{(}f^{\xi,y_{1}}(q),-f^{\xi,y_{2}}(q)\big{)}=\varnothing.

In case xx is a chain-chain singularity (S7), one may again follow the above reasoning to deduce that the existence of a connected component VV of EεcE_{\varepsilon}^{c} with xVx\in\partial V would contradict the existence of a sequence sn0s_{n}\to 0 with fξi,y1(sn)=fξi,y2(sn)f^{\xi_{i},y_{1}}(s_{n})=-f^{\xi_{i},y_{2}}(s_{n}), which is on the other hand guaranteed for both ξ1,ξ2Ξxext(Eε)\xi_{1},\xi_{2}\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) by Proposition 4.2. ∎

Remark 4.4.

Corollary 4.3 states that it is impossible for a chain singularity (S6) or a chain-chain singularity (S7) xx to lie on the boundary of any connected component VEεcV\subset E_{\varepsilon}^{c}, even though a sequence (Vn)n=1(V_{n})_{n=1}^{\infty} of connected components VnEεcV_{n}\subset E_{\varepsilon}^{c} converges to xx in Hausdorff distance. This is the motivation for the terminology of inaccessible singularities and can be seen as an analogue of the distinction between accessible and inaccessible points in a Cantor set C[0,1]C\subset[0,1], where inaccessible points do not lie on the boundary of any of the countably many removed open intervals [an,bn][0,1][a_{n},b_{n}]\subset[0,1]. See also Example 5.7 and the discussion after Definition 4.1 above.

Proposition 4.5 (Characterisation of chain singularities).

Let E2E\subset\mathbb{R}^{2}, let xEεx\in\partial E_{\varepsilon} and let 𝒢(x)\mathcal{G}(x) be a local boundary representation at xx with the functions gξ,y𝒢(x)g_{\xi,y}\in\mathcal{G}(x) as in (3.8). Then the following are equivalent:

  • (i)

    xx is a chain singularity (type S6, S7 or S8).

  • (ii)

    There exists a sequence (Vn)n=1(V_{n})_{n=1}^{\infty} of mutually disjoint connected components VnEεcV_{n}\subset E_{\varepsilon}^{c} for which distH(x,Vn)0\mathrm{dist}_{H}(x,V_{n})\to 0 as nn\to\infty.

  • (iii)

    There exists a sequence (xn)n=1(x_{n})_{n=1}^{\infty} of singularities on Eε\partial E_{\varepsilon} for which xnxx_{n}\to x and

    (4.7) limnyn(1)xnε,yn(2)xnε=1,\lim_{n\to\infty}\left\langle\frac{y_{n}^{(1)}-x_{n}}{\varepsilon},\frac{y_{n}^{(2)}-x_{n}}{\varepsilon}\right\rangle=-1,

    where ΠEext(xn)={yn(1),yn(2)}\Pi_{E}^{\mathrm{ext}}(x_{n})=\big{\{}y_{n}^{(1)},y_{n}^{(2)}\big{\}} for each nn\in\mathbb{N}.

  • (iv)

    The extremal contributors ΠEext(x)={y1,y2}\Pi_{E}^{\mathrm{ext}}(x)=\{y_{1},y_{2}\} satisfy y1x=(y2x)y_{1}-x=-(y_{2}-x) and there exist some ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) and corresponding functions gξ,y1,gξ,y2𝒢(x)g_{\xi,y_{1}},g_{\xi,y_{2}}\in\mathcal{G}(x) for which gξ,y1(sn)=gξ,y2(sn)g_{\xi,y_{1}}(s_{n})=g_{\xi,y_{2}}(s_{n}) for all nn\in\mathbb{N} for some sequence (sn)n=1+(s_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} with sn0s_{n}\to 0 as nn\to\infty.

Proof.

We begin by showing that (ii) \Rightarrow (iv) \Rightarrow (iii) \Rightarrow (ii). Since clearly (i) implies (iii), the result then follows by showing that (iii) \wedge (iv) \Rightarrow (i).

(ii) \Rightarrow (iv). Assume there exists a sequence (Vn)n=1Eεc(V_{n})_{n=1}^{\infty}\subset E_{\varepsilon}^{c} of mutually disjoint connected components of the complement EεcE_{\varepsilon}^{c} with distH(x,Vn)0\mathrm{dist}_{H}(x,V_{n})\to 0 as nn\to\infty. It follows then from Proposition 3.9 that xUnpε(E)x\notin\mathrm{Unp_{\varepsilon}}(E) and it cannot be a wedge, which implies y1x=(y2x)y_{1}-x=-(y_{2}-x) for the extremal contributors ΠEext(x)={y1,y2}\Pi_{E}^{\mathrm{ext}}(x)=\{y_{1},y_{2}\}.

For each nn\in\mathbb{N}, choose a point vnVnv_{n}\in V_{n}, and define ξn:=(vnx)/vnx\xi_{n}:=(v_{n}-x)/\left\lVert v_{n}-x\right\rVert. Due to compactness, one can choose a subsequence (vnk)k=1(v_{n_{k}})_{k=1}^{\infty} for which ξnkξS1\xi_{n_{k}}\to\xi\in S^{1}. Since vnEεcv_{n}\in E_{\varepsilon}^{c} for all nn\in\mathbb{N}, it follows from Lemma 2.9 that yix,ξ=0\langle y_{i}-x,\xi\rangle=0 for i{1,2}i\in\{1,2\}, so that ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}).

Assume contrary to the claim that there exists δ>0\delta>0 for which gξ,y1(s)gξ,y2(s)g_{\xi,y_{1}}(s)\neq g_{\xi,y_{2}}(s) for all s(0,δ)s\in(0,\delta). According to Proposition 4.2 this implies

(4.8) EεcUδ(x,ξ)=VUδ(x,ξ)=0<ρ<δρ(αρ,βρ)S1,E_{\varepsilon}^{c}\cap U_{\delta}(x,\xi)=V\cap U_{\delta}(x,\xi)=\bigcup_{0<\rho<\delta}\rho(\alpha_{\rho},\beta_{\rho})_{S^{1}},

where VV is the unique connected component of the complement EεcE_{\varepsilon}^{c} that intersects Uδ(x,ξ)U_{\delta}(x,\xi). Since ξnkξ\xi_{n_{k}}\to\xi as kk\to\infty, equation (4.8) now implies vnkVv_{n_{k}}\in V for large kk\in\mathbb{N}, which in turn contradicts the assumption that the sets VnV_{n} are connected and mutually disjoint. Hence, no such δ\delta can exist, and (iv) follows.

(iv) \Rightarrow (iii). Case (ii) in Proposition 4.2 now holds true, and directly implies (iii) here.

(iii) \Rightarrow (ii). Write ΠEext(x)={y1,y2}\Pi_{E}^{\mathrm{ext}}(x)=\{y_{1},y_{2}\} and define ξn:=(xnx)/xnx\xi_{n}:=(x_{n}-x)/\left\lVert x_{n}-x\right\rVert. Due to Lemma 2.13 (i) there exists a subsequence (xnk)k=1(x_{n_{k}})_{k=1}^{\infty}, for which ξnkξΞxext(Eε)\xi_{n_{k}}\to\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) as kk\to\infty. Furthermore, since ΠEext(xn)={yn(1),yn(2)}\Pi_{E}^{\mathrm{ext}}(x_{n})=\big{\{}y_{n}^{(1)},y_{n}^{(2)}\big{\}} for all nn\in\mathbb{N}, Lemma 2.13 (ii)(a) together with (4.7) implies the existence of a further subsequence (xm)m=1(xnk)k=1(x_{m})_{m=1}^{\infty}\subset(x_{n_{k}})_{k=1}^{\infty} for which ym(i)yiΠEext(x)y_{m}^{(i)}\to y_{i}\in\Pi_{E}^{\mathrm{ext}}(x) for i{1,2}i\in\{1,2\}. Hence y1x=(y2x)y_{1}-x=-(y_{2}-x).

To complete the argument we show that for any δ>0\delta>0 there exists some 0<sm<δ0<s_{m}<\delta for which xm=gξ,y1(sm)=gξ,y2(sm)x_{m}=g_{\xi,y_{1}}(s_{m})=g_{\xi,y_{2}}(s_{m}). Once this is established, the statement follows from Proposition 4.2. Since xmxx_{m}\to x as mm\to\infty, there exists for all δ>0\delta>0 some MM\in\mathbb{N} for which xmxδ\left\lVert x_{m}-x\right\rVert\leq\delta for all m>Mm>M. Hence there exists for all m>Mm>M some smδs_{m}\leq\delta for which xm=gξ,yi(sm)x_{m}=g_{\xi,y_{i}}(s_{m}) for some i{1,2}i\in\{1,2\}. But since ym(i)yiΠEext(x)y_{m}^{(i)}\to y_{i}\in\Pi_{E}^{\mathrm{ext}}(x) for i{1,2}i\in\{1,2\} as mm\to\infty, where y1x=(y2x)y_{1}-x=-(y_{2}-x), we have in fact xm=gξ,y1(sm)=gξ,y2(sm)x_{m}=g_{\xi,y_{1}}(s_{m})=g_{\xi,y_{2}}(s_{m}) for all sufficiently large m>Mm>M.

(iii) \wedge (iv) \Rightarrow (i). Since y1x=(y2x)y_{1}-x=-(y_{2}-x) for the extremal contributors y1,y2ΠEext(x)y_{1},y_{2}\in\Pi_{E}^{\mathrm{ext}}(x), it follows from Lemma 2.10 that the extremal outward directions ξ1,ξ2Ξxext(Eε)\xi_{1},\xi_{2}\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) satisfy either ξ1=ξ2\xi_{1}=\xi_{2} or ξ1=ξ2\xi_{1}=-\xi_{2}. In the former case xx is a one-sided chain singularity (S6). In the latter case we may assume without loss of generality that ξn:=(xnx)/xnxξ1\xi_{n}:=(x_{n}-x)/\left\lVert x_{n}-x\right\rVert\to\xi_{1}. Proposition 4.2 then states that for some δ>0\delta>0, the boundary subset EεUδ(x,ξ2)\partial E_{\varepsilon}\cap U_{\delta}(x,\xi_{2}) exhibits either 'sharp'-type or 'chain'-type geometry and that these cases are mutually exclusive. In the former case xx is a sharp-chain singularity (S8), in the latter a chain-chain singularity (S7). ∎

We employ Proposition 4.5 to show that our definition of smooth points (see Definition 2.2) coincides with the property of lying on a C1C^{1}-smooth curve. By a curve we mean the image Γ=γ([0,1])\Gamma=\gamma([0,1]) of a continuous, injective map γ:[0,1]2\gamma:[0,1]\to\mathbb{R}^{2}.

Proposition 4.6 (Characterisation of smooth points).

Let E2E\subset\mathbb{R}^{2} and xEεx\in\partial E_{\varepsilon}. Then xx is smooth in the sense of Definition 2.2 if and only if there exists a C1C^{1}-curve Γ\Gamma for which Γ=EεBr(x)¯\Gamma=\partial E_{\varepsilon}\cap\overline{B_{r}(x)} for some δ>0\delta>0.

Proof.

According to Proposition 3.5 there exists a local boundary representation 𝒢(x)\mathcal{G}(x) with radius r>0r>0 and continuous functions fξ,y:[0,ε/2]f^{\xi,y}:[0,\varepsilon/2]\to\mathbb{R} for which

gξ,y(s)=x+sξ+fξ,y(s)(xy)g_{\xi,y}(s)=x+s\xi+f^{\xi,y}(s)(x-y)

for all gξ,y𝒢(x)g_{\xi,y}\in\mathcal{G}(x) and s[0,sξ,y]s\in[0,s_{\xi,y}].

(i) Assume first that xx is smooth in the sense of Definition 2.2. Then there exists 0<δ<r0<\delta<r for which Bδ(x)¯EεUnpε(E)\overline{B_{\delta}(x)}\cap\partial E_{\varepsilon}\subset\mathrm{Unp_{\varepsilon}}(E). Proposition 2.12 then implies that for all zBδ(x)¯Eεz\in\overline{B_{\delta}(x)}\cap\partial E_{\varepsilon} the set of extremal outward directions satisfies Ξzext(Eε)={ξ,ξ}\Xi_{z}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi,-\xi\} for some ξS1\xi\in S^{1}. According to Proposition 2.14, the extremal outward directions coincide with tangential directions on the boundary, and hence Ξzext(Eε)={ξ,ξ}\Xi_{z}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi,-\xi\} implies that gξ,yg_{\xi,y} is differentiable at s=szs=s_{z}, where z=gξ,y(sz)z=g_{\xi,y}(s_{z}). Since this is true for all zBδ(x)¯Eεz\in\overline{B_{\delta}(x)}\cap\partial E_{\varepsilon}, the boundary Eε\partial E_{\varepsilon} inside Bδ(x)¯\overline{B_{\delta}(x)} is contained in the union of images gξ1,y([0,sξ1,y])gξ2,y([0,sξ2,y])g_{\xi_{1},y}([0,s_{\xi_{1},y}])\cup g_{\xi_{2},y}([0,s_{\xi_{2},y}]) under the differentiable maps gξ1,yg_{\xi_{1},y} and gξ2,yg_{\xi_{2},y}. Since gξ1,y([0,sξ1,y])gξ2,y([0,sξ2,y])={x}g_{\xi_{1},y}([0,s_{\xi_{1},y}])\cap g_{\xi_{2},y}([0,s_{\xi_{2},y}])=\{x\} and both images can be represented as graphs of the corresponding functions fξi,yf^{\xi_{i},y} for i{1,2}i\in\{1,2\}, the claim follows.

(ii) Let then Γ\Gamma be a C1C^{1}-curve for which Γ=EεBδ(x)¯\Gamma=\partial E_{\varepsilon}\cap\overline{B_{\delta}(x)} for some δ>0\delta>0. Then for each zEεBδ(x)¯z\in\partial E_{\varepsilon}\cap\overline{B_{\delta}(x)}, Proposition 2.14 implies Ξzext(Eε)={ξ,ξ}\Xi_{z}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi,-\xi\} for some ξS1\xi\in S^{1}. Consider now some zEεBδ(x)¯z\in\partial E_{\varepsilon}\cap\overline{B_{\delta}(x)}. Since Γ\Gamma is C1C^{1}-smooth, the correspondence between tangents and extremal outward directions given by Proposition 2.14 implies that zz is not a wedge (S1) or a sharp singularity (S2–S3). On the other hand, as a curve Γ\Gamma is connected, which together with Proposition 4.5 implies that zz cannot be a chain singularity (S6–S8). Hence it follows from Theorem 1 below that zz is necessarily either a smooth point or a shallow (S4–S5) singularity, which implies zUnpε(E)z\in\mathrm{Unp_{\varepsilon}}(E). The same argument applies to all zEεBr(x)¯z\in\partial E_{\varepsilon}\cap\overline{B_{r}(x)}, which means that xx is smooth in the sense of Definition 2.2. ∎

4.2.1. Proof of Theorem 1

We conclude this section with the proof of our first main result, a classification of boundary points on Eε\partial E_{\varepsilon}. We restate the result here for the convenience of the reader.

Theorem 1 (Classification of boundary points).

Let E2E\subset\mathbb{R}^{2} be compact, ε>0\varepsilon>0 and let xEεx\in\partial E_{\varepsilon} be a boundary point of EεE_{\varepsilon} that is not smooth. Then xx belongs to precisely one of the eight categories of singularities given in Definition 4.1.

Proof.

Note first that for any u,vS1u,v\in S^{1} either u=vu=v, u=vu=-v, or u{v,v}u\notin\{v,-v\}. Hence we obtain the following categorisation of boundary point types according to the orientation of the extremal outward directions Ξxext(Eε)={ξ1,ξ2}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi_{1},\xi_{2}\}:

ξ1\displaystyle\xi_{1} {ξ2,ξ2}:\displaystyle\notin\{\xi_{2},-\xi_{2}\}: type S1
ξ1\displaystyle\xi_{1} =ξ2:\displaystyle=\xi_{2}: types S2 and S6
ξ1\displaystyle\xi_{1} =ξ2:\displaystyle=-\xi_{2}: smooth points and types S3–S5, S7–S8

These are due to Proposition 2.12 for xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E) and Definition 4.1 for xUnpε(E)x\notin\mathrm{Unp_{\varepsilon}}(E), and they correspond to the cases (a)–(c) illustrated in Figure 9. It follows immediately that if xx is a wedge (S1), it cannot be of any other type, and vice versa. In addition, of all the defined types of boundary points, only the shallow singularities (S4 and S5) and smooth points satisfy ΠEext(x)={y}\Pi_{E}^{\mathrm{ext}}(x)=\{y\} for some yEy\in\partial E, and these types are by definition mutually exclusive.

Hence it suffices to show that the remaining types S2–S3 and S6–S8 (corresponding to case (c) in Figure 9) are mutually exclusive. For all these types, the set of extremal contributors ΠEext(x)={y1,y2}\Pi_{E}^{\mathrm{ext}}(x)=\{y_{1},y_{2}\} satisfies y1x=(y2x)y_{1}-x=-(y_{2}-x). Proposition 4.2 then states that for each ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) either

  • (i)

    there exists a connected component VEεcV\subset E_{\varepsilon}^{c} and r>0r>0 for which EεcUr(x,ξ)=VUr(x,ξ)E_{\varepsilon}^{c}\cap U_{r}(x,\xi)=V\cap U_{r}(x,\xi) and xVx\in\partial V, or

  • (ii)

    there exists a sequence of singularities (xn)n=1(x_{n})_{n=1}^{\infty} in S(Eε)S(E_{\varepsilon}) with xnxx_{n}\to x as nn\to\infty and (xnx)/xnxξ(x_{n}-x)/\left\lVert x_{n}-x\right\rVert\to\xi, and

    limnyn(1)xnε,yn(2)xnε=1,\lim_{n\to\infty}\left\langle\frac{y_{n}^{(1)}-x_{n}}{\varepsilon},\frac{y_{n}^{(2)}-x_{n}}{\varepsilon}\right\rangle=-1,

and that these situations are mutually exclusive. In other words, for each extremal outward direction ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}), the intersection EεUr(x,ξ)\partial E_{\varepsilon}\cap U_{r}(x,\xi) exhibits either 'sharp'-type or 'chain'-type geometry (see Definition 4.1). In case ξ1=ξ2\xi_{1}=\xi_{2}, the point xx is hence either a sharp (S2) or a chain (S6) singularity, and in case ξ2=ξ1\xi_{2}=-\xi_{1}, it is either a sharp-sharp (S3), a chain-chain (S7), or a sharp-chain (S8) singularity, and all these cases are mutually exclusive. ∎

5. Topological Structure of the Set of Singularities

Since the categories of boundary points given in Definition 4.1 define a partition of the boundary, it makes sense to inquire on their cardinalities and topological structure. Our second main result, Theorem 2, states that for any compact E2E\in\mathbb{R}^{2} and ε>0\varepsilon>0, the sets of wedges (S1), sharp singularities (S2, S3 and S8) and one-sided chain singularities (S6) on Eε\partial E_{\varepsilon} are at most countably infinite. This does not hold in general for the sets of shallow-shallow singularities (S5) or chain-chain singularities (S7), which may even have a positive one-dimensional Hausdorff measure on the boundary (see [14, 15]). In Section 5.2 we show that the set 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) of chain singularities is nevertheless nowhere dense, and hence small in the topological sense.

5.1. Cardinalities of Sets of Singularities

In order to prove the above-mentioned results on the cardinalities of the sets of singularities, we proceed by treating one by one the cases of wedges, sharp singularities, and one-sided shallow and chain singularities. We begin with the following general result on the geometry of accumulating singularities, which is essentially a corollary of Lemma 2.13 on the asymptotic behaviour of sequences of boundary points.

Lemma 5.1 (Geometry of accumulating singularities).

Let E2E\subset\mathbb{R}^{2}, let (xn)n=1S(Eε)(x_{n})_{n=1}^{\infty}\subset S(E_{\varepsilon}) be a sequence of pair-wise disjoint singularities with xnxEεx_{n}\to x\in\partial E_{\varepsilon} and let Ξxnext(Eε)={ξn(1),ξn(2)}\Xi_{x_{n}}^{\mathrm{ext}}(E_{\varepsilon})=\big{\{}\xi_{n}^{(1)},\xi_{n}^{(2)}\big{\}} for each nn\in\mathbb{N}. Then limn|ξn(1),ξn(2)|=1\lim_{n\to\infty}\big{|}\big{\langle}\xi_{n}^{(1)},\xi_{n}^{(2)}\big{\rangle}\big{|}=1.

Proof.

Due to Lemma 2.13 (i) we can assume that the limit ξ:=limn(xnx)/xnx\xi:=\lim_{n\to\infty}(x_{n}-x)/\left\lVert x_{n}-x\right\rVert exists. Note that depending on the geometry at the singularities xnx_{n}, each of the extremal outward directions ξn(i)\xi_{n}^{(i)} for i{1,2}i\in\{1,2\} may be more aligned with ξ\xi than ξ-\xi, or vice versa. However, for each i{1,2}i\in\{1,2\}, Lemma 2.13 (ii)(b) implies that there exist sequences of coefficients (an(i))n=1{1,1}\big{(}a_{n}^{(i)}\big{)}_{n=1}^{\infty}\in\{1,-1\}^{\mathbb{N}} for which |ξn(i),ξ|=an(i)ξn(i),ξ\big{|}\big{\langle}\xi_{n}^{(i)},\xi\big{\rangle}\big{|}=\big{\langle}a_{n}^{(i)}\xi_{n}^{(i)},\xi\big{\rangle} and

(5.1) limnan(1)ξn(1),ξ=limnan(2)ξn(2),ξ=1.\lim_{n\to\infty}\big{\langle}a_{n}^{(1)}\xi_{n}^{(1)},\xi\big{\rangle}=\lim_{n\to\infty}\big{\langle}a_{n}^{(2)}\xi_{n}^{(2)},\xi\big{\rangle}=1.

Let θn(i)0\theta_{n}^{(i)}\geq 0 be the angle for which an(i)ξn(i),ξ=cos(θn(i))\big{\langle}a_{n}^{(i)}\xi_{n}^{(i)},\xi\big{\rangle}=\cos\big{(}\theta_{n}^{(i)}\big{)}. Equation (5.1) implies θn(i)0\theta_{n}^{(i)}\to 0 for i{1,2}i\in\{1,2\}, and in particular there exists some NN\in\mathbb{N} for which |ξn(1),ξn(2)|=an(1)ξn(1),an(2)ξn(2)\big{|}\big{\langle}\xi_{n}^{(1)},\xi_{n}^{(2)}\big{\rangle}\big{|}=\big{\langle}a_{n}^{(1)}\xi_{n}^{(1)},a_{n}^{(2)}\xi_{n}^{(2)}\big{\rangle} for all nNn\geq N. Writing θnmax:=max{θn(1),θn(2)}\theta_{n}^{\textrm{max}}:=\max\big{\{}\theta_{n}^{(1)},\theta_{n}^{(2)}\big{\}} for nn\in\mathbb{N}, we have

(5.2) an(1)ξn(1),an(2)ξn(2)=cos(θn(1)+θn(2))cos(2θnmax)1,\big{\langle}a_{n}^{(1)}\xi_{n}^{(1)},a_{n}^{(2)}\xi_{n}^{(2)}\big{\rangle}=\cos\big{(}\theta_{n}^{(1)}+\theta_{n}^{(2)}\big{)}\geq\cos\left(2\theta_{n}^{\textrm{max}}\right)\longrightarrow 1,

as nn\to\infty, and the result follows. ∎

A particular consequence of Lemma 5.1 is that accumulating wedges (S1) become increasingly acute/obtuse as they approach a limit point, with the angles θn:=(ξn(1),ξn(2))\theta_{n}:=\sphericalangle\big{(}\xi_{n}^{(1)},\xi_{n}^{(2)}\big{)} between the extremal outward directions approaching an asymptotic value θ=limnθn{0,π}\theta=\lim_{n\to\infty}\theta_{n}\in\{0,\pi\}. It follows that for any fixed p>0p>0, there can only exist finitely many wedges whose sharpness deviates from these asymptotic values by more than pp, which in turn implies that the total number of wedges can at most be countably infinite. Lemma 5.2 below makes this argument precise.

Lemma 5.2 (Number of wedges).

For any compact E2E\subset\mathbb{R}^{2} and ε>0\varepsilon>0, the number of wedges (S1) on Eε\partial E_{\varepsilon} is at most countably infinite.

Proof.

We begin by showing that for each p(0,1)p\in(0,1), the subset A(p)EεA(p)\subset\partial E_{\varepsilon} defined by

A(p):={xEε:|ξ(1),ξ(2)|pforξ(1),ξ(2)Ξxext(Eε)}A(p):=\big{\{}x\in\partial E_{\varepsilon}\,:\,\big{|}\big{\langle}\xi^{(1)},\xi^{(2)}\big{\rangle}\big{|}\leq p\,\,\,\,\textrm{for}\,\,\,\,\xi^{(1)},\xi^{(2)}\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})\big{\}}

contains only finitely many points. Assume contrary to this that for some p(0,1)p\in(0,1) the set A(p)A(p) contains infinitely many points. This implies that there exists a pair-wise disjoint sequence (xn)n=1(x_{n})_{n=1}^{\infty} with xnA(p)x_{n}\in A(p) for all nn\in\mathbb{N}, and due to compactness of Eε\partial E_{\varepsilon} we may assume that xnxEεx_{n}\to x\in\partial E_{\varepsilon} as nn\to\infty. Writing Ξxnext(Eε)={ξn(1),ξn(2)}\Xi_{x_{n}}^{\mathrm{ext}}(E_{\varepsilon})=\big{\{}\xi_{n}^{(1)},\xi_{n}^{(2)}\big{\}} for each nn\in\mathbb{N}, Lemma 5.1 then implies limn|ξn(1),ξn(2)|=1\lim_{n\to\infty}\big{|}\big{\langle}\xi_{n}^{(1)},\xi_{n}^{(2)}\big{\rangle}\big{|}=1, contradicting the assumption that |ξn(1),ξn(2)|p<1\big{|}\big{\langle}\xi_{n}^{(1)},\xi_{n}^{(2)}\big{\rangle}\big{|}\leq p<1 for all nn\in\mathbb{N}.

By definition, each wedge xEεx\in\partial E_{\varepsilon} belongs to the set A(p)A(p) for some p(0,1)p\in(0,1). Hence the union

A:=n=1A(1n1)A:=\bigcup_{n=1}^{\infty}A\big{(}1-n^{-1}\big{)}

contains all the wedges on Eε\partial E_{\varepsilon}. According to the reasoning above, each of the sets A(1n1)A(1-n^{-1}) contains only finitely many points, from which the result follows. ∎

We next show that for a given connected component UU of the complement EεcE_{\varepsilon}^{c}, there can only exist finitely many sharp singularities on U\partial U. This essentially follows from Propositions 3.5 and 4.5 which imply that any convergent sequence of pairwise disjoint sharp singularities xnUx_{n}\in\partial U is associated with a sequence (Un)n=1(U_{n})_{n=1}^{\infty} of pairwise disjoint connected components of EεcE_{\varepsilon}^{c}, aligned with one of the extremal outward directions ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) and satisfying UnUU_{n}\neq U and xnUnx_{n}\in\partial U_{n} for all nn\in\mathbb{N} (or, to be precise, at least from some NN\in\mathbb{N} onwards). Negotiating the topological and geometric constraints of the situation would then require that U=UnU=U_{n} for all nn\in\mathbb{N}, which is clearly impossible.

Lemma 5.3 (Number of sharp singularities).

Let E2E\subset\mathbb{R}^{2} be compact. For any connected component UU of the complement EεcE_{\varepsilon}^{c}, the number of sharp singularities (S2, S3 and S8) on the boundary U\partial U is finite.

Proof.

Assume that the claim fails for some connected UEεcU\subset E_{\varepsilon}^{c}. Since U\partial U is compact, this implies the existence of a pair-wise disjoint sequence (xn)n=1U(x_{n})_{n=1}^{\infty}\subset\partial U of sharp singularities with xnxUx_{n}\to x\in\partial U. We may furthermore assume that the sequence is ordered so that

(5.3) xn+1x<xnx\left\lVert x_{n+1}-x\right\rVert<\left\lVert x_{n}-x\right\rVert

for all nn\in\mathbb{N}, and that the limit ξ:=limn(xnx)/xnx\xi:=\lim_{n\to\infty}(x_{n}-x)/\left\lVert x_{n}-x\right\rVert exists. Write ΠEext(xn)={yn(1),yn(2)}\Pi_{E}^{\mathrm{ext}}(x_{n})=\big{\{}y_{n}^{(1)},y_{n}^{(2)}\big{\}}. It follows from the definition of a sharp singularity that

(5.4) yn(1)xn=(yn(2)xn)y_{n}^{(1)}-x_{n}=-\big{(}y_{n}^{(2)}-x_{n}\big{)}

for all nn\in\mathbb{N}. According to Proposition 4.5, also the extremal contributors y1,y2ΠEext(x)y_{1},y_{2}\in\Pi_{E}^{\mathrm{ext}}(x) satisfy y1x=(y2x)y_{1}-x=-(y_{2}-x), and yn(i)yiy_{n}^{(i)}\to y_{i} for i{1,2}i\in\{1,2\} as nn\to\infty. Let 𝒢(x)\mathcal{G}(x) be the local boundary representation (of radius r>0r>0) at xx, given by Proposition 3.5, and consider for i{1,2}i\in\{1,2\} the functions gξ,yi𝒢(x)g_{\xi,y_{i}}\in\mathcal{G}(x),

gξ,yi(s):=x+sξ+fξ,yi(s)(xyi)g_{\xi,y_{i}}(s):=x+s\xi+f^{\xi,y_{i}}(s)(x-y_{i})

where the functions fξ,yi:[0,ε/2]f^{\xi,y_{i}}:[0,\varepsilon/2]\to\mathbb{R} are continuous. Due to equation (5.4), and since xnxx_{n}\to x, there exists a sequence (sn)n=1+(s_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} with sn0s_{n}\to 0 and some NN\in\mathbb{N}, for which n>Nn>N implies xn=gξ,y1(sn)=gξ,y2(sn)x_{n}=g_{\xi,y_{1}}(s_{n})=g_{\xi,y_{2}}(s_{n}) and consequently fξ,y1(sn)+fξ,y2(sn)=0f^{\xi,y_{1}}(s_{n})+f^{\xi,y_{2}}(s_{n})=0. Inequality (5.3) implies sn+1<sn<sn1s_{n+1}<s_{n}<s_{n-1} for all n>Nn>N, and one can define the open sets

Sn\displaystyle S_{n} :={s(sn+1,sn1):fξ,y1(s)+fξ,y2(s)<0},\displaystyle:=\left\{s\in(s_{n+1},s_{n-1})\,:\,f^{\xi,y_{1}}(s)+f^{\xi,y_{2}}(s)<0\right\},
Un\displaystyle U_{n} :={τgξ,y1(s)+(1τ)gξ,y2(s):τ(0,1),sSn}.\displaystyle:=\left\{\tau g_{\xi,y_{1}}(s)+(1-\tau)g_{\xi,y_{2}}(s)\,:\,\tau\in(0,1),\,s\in S_{n}\right\}.

For each n>Nn>N, the set UnU_{n} is contained in the interior intRn\textrm{int}\,R_{n} of the closed rectangle

Rn:={x+sξ+t(xy1):sn+1ssn1,infsSn{fξ,y1(s)}tinfsSn{fξ,y2(s)}}.R_{n}:=\Big{\{}x+s\xi+t(x-y_{1})\,:\,s_{n+1}\leq s\leq s_{n-1},\,\inf_{s\in S_{n}}\left\{f^{\xi,y_{1}}(s)\right\}\leq t\leq-\inf_{s\in S_{n}}\left\{f^{\xi,y_{2}}(s)\right\}\Big{\}}.

By definition snSns_{n}\in S_{n} for all n>Nn>N, from which it follows that xnVx_{n}\notin\partial V for any open VEεcRncV\subset E_{\varepsilon}^{c}\cap R_{n}^{c}. On the other hand xnUx_{n}\in\partial U for all nn\in\mathbb{N} which implies UintRnU\subset\textrm{int}\,R_{n} for all nn\in\mathbb{N}, since UU is connected and RnEεc=\partial R_{n}\cap E_{\varepsilon}^{c}=\varnothing by definition. However, given that xURncx\in\partial U\cap R_{n}^{c}, this leads to the contradiction URnRncU\subset R_{n}\cap R_{n}^{c} for all n>Nn>N. ∎

Lemmas 5.4 and 5.6 below state that the sets of one-sided shallow singularities (S4) and chain singularities (S6) are both at most countably infinite. The argument in both cases rests on the fact that for any finite sum M:=xAmx<M:=\sum_{x\in A}m_{x}<\infty of non-negative real numbers mxm_{x} indexed by a (potentially uncountable) set AA, the index subset A0:={xA:mx>0}A_{0}:=\{x\in A\,:\,m_{x}>0\} corresponding to the positive elements in the sum is at most countably infinite.222This follows from the observation that the set An:={xA:mx>1/n}A_{n}:=\{x\in A\,:\,m_{x}>1/n\} is finite for each nn\in\mathbb{N} and hence A0:={xA:mx>0}=nAnA_{0}:=\{x\in A\,:\,m_{x}>0\}=\bigcup_{n\in\mathbb{N}}A_{n} is countable as a countable union of finite sets. In the case of shallow singularities, the numbers being summed will represent lengths (one-dimensional Hausdorff measures) mx:=1(Ix)m_{x}:=\mathcal{H}^{1}(I_{x}) of boundary segments IxEεI_{x}\subset\partial E_{\varepsilon}, and in the case of chain singularities they will stand for surface areas mx:=2(Ax)m_{x}:=\mathcal{H}^{2}(A_{x}) of open subsets AxintEεA_{x}\subset\mathrm{int}\,E_{\varepsilon}. In each case, these numbers will be strictly positive by definition for every xx, implying that the underlying index sets—corresponding to the sets of singularities in question—are themselves at most countably infinite.

Lemma 5.4 (Number of one-sided shallow singularities).

For a compact set E2E\subset\mathbb{R}^{2}, the number of one-sided shallow singularities (S4) on Eε\partial E_{\varepsilon} is at most countably infinite.

Proof.

Write WW for the set of one-sided shallow singularities on Eε\partial E_{\varepsilon}, and consider some xWx\in W. By definition of a one-sided shallow singularity, there exists a ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) and δ>0\delta>0 for which Jx:=Uδ(x,ξ)EεUnpε(E)J_{x}:=U_{\delta}(x,\xi)\cap\partial E_{\varepsilon}\subset\mathrm{Unp_{\varepsilon}}(E). Due to Proposition 3.5 there exists a local boundary representation 𝒢(x)\mathcal{G}(x) and some some r(x)>0r(x)>0 for which Jx=gξ,y([0,r(x)])J_{x}=g_{\xi,y}([0,r(x)]), where ΠEext(x)={y}\Pi_{E}^{\mathrm{ext}}(x)=\{y\} and gξ,y𝒢(x)g_{\xi,y}\in\mathcal{G}(x), satisfies

gξ,y(s)=x+sξ+fξ,y(s)(xy)g_{\xi,y}(s)=x+s\xi+f^{\xi,y}(s)(x-y)

for some continuous f:[0,ε/2]f:[0,\varepsilon/2]\to\mathbb{R}. In particular, the open subsegment Ix:=gξ,y((0,r(x)/2))JxI_{x}:=g_{\xi,y}((0,r(x)/2))\subset J_{x} contains only smooth points and has a positive, finite length r(x)/2<1(Ix)<r(x)/2<\mathcal{H}^{1}(I_{x})<\infty. This follows for instance from the fact that gξ,yg_{\xi,y} satisfies |sw|gξ,y(s)gξ,y(w)2|sw|/3|s-w|\leq\left\lVert g_{\xi,y}(s)-g_{\xi,y}(w)\right\rVert\leq 2|s-w|/\sqrt{3} for all s,w(0,r(x)/2)s,w\in(0,r(x)/2) (see Proposition 3.6) and since the increase of the Hausdorff measure under a Lipschitz map is bounded by the Lipschitz constant (see for instance [1, Proposition 2.49]). If zxz\neq x is another one-sided shallow singularity, the corresponding segment IzI_{z} satisfies IzIx=I_{z}\cap I_{x}=\varnothing by definition. Hence, the collection I:=xWIxEεI:=\bigcup_{x\in W}I_{x}\subset\partial E_{\varepsilon} has finite length

(5.5) 1(I)=xW1(Ix)<1(Eε)<.\mathcal{H}^{1}(I)=\sum_{x\in W}\mathcal{H}^{1}(I_{x})<\mathcal{H}^{1}(E_{\varepsilon})<\infty.

The last inequality in (5.5) was established already by Erdős in [11, Section 6]. According to the counting argument preceding the statement of the result, inequality (5.5) implies that the set WW can be at most countably infinite. ∎

The following example demonstrates that the set of two-sided shallow singularities (S5) can be dense and have positive Hausdorff measure on the boundary Eε\partial E_{\varepsilon}. The idea is to construct a suitably jagged function on the interval [0,1][0,1] (say) and interpret its graph as a subset of the boundary Eε\partial E_{\varepsilon} of a corresponding set E2E\subset\mathbb{R}^{2}.

Example 5.5 (Dense, positive measure set of shallow singularities).

Consider a bounded, increasing function α:[0,1]\alpha:[0,1]\to\mathbb{R} that is discontinuous at every rational number q[0,1]q\in\mathbb{Q}\cap[0,1] but continuous at every irrational number p[0,1]p\in[0,1]\setminus\mathbb{Q}.333One way to construct such a function is to write [0,1]={qn:n}\mathbb{Q}\cap[0,1]=\big{\{}q_{n}\,:\,n\in\mathbb{N}\big{\}}, define N(x):={n:qnx}N(x):=\{n\,:\,q_{n}\leq x\}, take any positive summable sequence (an)n=1(a_{n})_{n=1}^{\infty}, and set α(x):=nN(x)an\alpha(x):=\sum_{n\in N(x)}a_{n} for all x[0,1]x\in[0,1]. This way α\alpha is increasing on [0,1][0,1], has a jump of amplitude ana_{n} at each rational x=qnx=q_{n}, is continuous at every x[0,1]x\in[0,1]\setminus\mathbb{Q}, and satisfies α(1)=n=1an<\alpha(1)=\sum_{n=1}^{\infty}a_{n}<\infty.  As an almost everywhere continuous bounded function, every such α\alpha is Riemann-integrable, and its monotonicity implies that the integral function Iα(x):=0xα(s)𝑑sI_{\alpha}(x):=\int_{0}^{x}\alpha(s)ds is convex. Most significantly for our example, IαI_{\alpha} has a well-defined derivative at every irrational p[0,1]p\in[0,1]\setminus\mathbb{Q}, but not at any rational q[0,1]q\in\mathbb{Q}\cap[0,1].

For any ε>0\varepsilon>0, one may thus interpret the graph G:={(s,Iα(s)):s[0,1]}G:=\big{\{}(s,I_{\alpha}(s))\,:\,s\in[0,1]\big{\}} as a subset of an boundary Eε\partial E_{\varepsilon} as follows. Since the one-sided derivatives

D±Iα(s):=limh±0Iα(s+h)Iα(s)hD^{\pm}I_{\alpha}(s):=\lim_{h\to\pm 0}\frac{I_{\alpha}(s+h)-I_{\alpha}(s)}{h}

exist at every s[0,1]s\in[0,1], one can define for each x(s):=(s,Iα(s))x(s):=(s,I_{\alpha}(s)) the corresponding contributors y(s),y+(s)ΠE(x(s))y^{-}(s),y^{+}(s)\in\Pi_{E}(x(s)) by setting y±(s):=(s+a±(s),Iα(s)b±(s))y^{\pm}(s):=\big{(}s+a^{\pm}(s),\,I_{\alpha}(s)-b^{\pm}(s)\big{)}, where for each s[0,1]s\in[0,1]

a±(s):=εD±Iα(s)1+[D±Iα(s)]2andb±(s):=ε1+[D±Iα(s)]2.a^{\pm}(s):=\frac{\varepsilon D^{\pm}I_{\alpha}(s)}{\sqrt{1+\big{[}D^{\pm}I_{\alpha}(s)\big{]}^{2}}}\qquad\textrm{and}\qquad b^{\pm}(s):=\frac{\varepsilon}{\sqrt{1+\big{[}D^{\pm}I_{\alpha}(s)\big{]}^{2}}}.

A direct computation shows that D±Iα(s)=b±(s)/a±(s)D^{\pm}I_{\alpha}(s)=b^{\pm}(s)/a^{\pm}(s) and x(s)y±(s)=ε\left\lVert x(s)-y^{\pm}(s)\right\rVert=\varepsilon for all s[0,1]s\in[0,1]. Furthermore, the convexity of IαI_{\alpha} implies that for each y±(s)y^{\pm}(s) the ball Bε(y±(s))¯\overline{B_{\varepsilon}(y^{\pm}(s))} intersects GG only at the corresponding x(s)x(s), which implies that GEεG\subset\partial E_{\varepsilon} for the set E:={y±(s):s[0,1]}E:=\{y^{\pm}(s)\,:\,s\in[0,1]\}.

By construction, y+(p)=y(p)y^{+}(p)=y^{-}(p) for the irrational p[0,1]p\in[0,1]\setminus\mathbb{Q}, whereas y+(q)>y(q)y^{+}(q)>y^{-}(q) for all rational q[0,1]q\in\mathbb{Q}\cap[0,1]. Hence x(q)x(q) is a wedge for each rational qq while x(p)Unpε(E)x(p)\in\mathrm{Unp_{\varepsilon}}(E) for every irrational pp, which implies that the points x(p)x(p) are in fact two-sided shallow singularities (S5). Due to the continuity of IαI_{\alpha}, these points form a dense set on GG. Given that IαI_{\alpha} is convex and absolutely continuous as an integral function, and the derivative α\alpha is bounded, IαI_{\alpha} is in fact Lipschitz continuous with some Lipschitz constant KK. It then follows from the basic properties of Hausdorff measure (see for instance [1, Proposition 2.49]) and the rectifiability of Eε\partial E_{\varepsilon} (see [27, Proposition 2.3] and [12, Corollary 3.3]) that 1([0,1])1(G)<\mathcal{H}^{1}([0,1])\leq\mathcal{H}^{1}(G)<\infty and the set P:={(p,Iα(p)):p[0,1]}P:=\big{\{}(p,I_{\alpha}(p))\,:\,p\in[0,1]\setminus\mathbb{Q}\big{\}} of two-sided singularities has full measure on GG.

Lemma 5.6 (Number of one-sided chain singularities).

For a compact set E2E\subset\mathbb{R}^{2}, the number of one-sided chain singularities (S6) on Eε\partial E_{\varepsilon} is at most countably infinite.

Proof.

Write CC for the set of one-sided chain singularities on Eε\partial E_{\varepsilon}. We argue that there exists a collection {Ax}xC\{A_{x}\}_{x\in C} of pair-wise disjoint open sets AxintEεA_{x}\subset\textrm{int}\,E_{\varepsilon}, indexed by CC. The result then follows from the counting argument discussed in the lead-up to Lemma 5.4 above.

For each xCx\in C, the set of outward directions is a singleton Ξx(Eε)={ξ}\Xi_{x}(E_{\varepsilon})=\{\xi\}. Let xCx\in C and let 𝒢(x)\mathcal{G}(x) be a local boundary representation at xx so that for i{1,2}i\in\{1,2\} and gξ,yi𝒢(x)g_{\xi,y_{i}}\in\mathcal{G}(x)

gξ,yi(s)=x+sξ+fξ,yi(s)(xyi)g_{\xi,y_{i}}(s)=x+s\xi+f^{\xi,y_{i}}(s)(x-y_{i})

for some continuous functions fξ,yi:[0,ε/2]f^{\xi,y_{i}}:[0,\varepsilon/2]\to\mathbb{R}. It follows from the definition of one-sided chain singularities that the extremal contributors y1,y2ΠEext(x)y_{1},y_{2}\in\Pi_{E}^{\mathrm{ext}}(x) satisfy y1x=(y2x)y_{1}-x=-(y_{2}-x) for all xCx\in C. Since ξΞx(Eε)-\xi\notin\Xi_{x}(E_{\varepsilon}), there are two possibilities:

  • (i)

    there exists some non-extremal contributor yΠE(x)y\in\Pi_{E}(x) for which yx,ξ<0\langle y-x,\xi\rangle<0, or else

  • (ii)

    there exists some δ0+\delta_{0}\in\mathbb{R}_{+} such that for all δ<δ0\delta<\delta_{0},

    Bε(EUδ(y1,ξ))¯Bε(EUδ(y2,ξ))¯{x}\overline{B_{\varepsilon}(E\cap U_{\delta}(y_{1},-\xi))}\cap\overline{B_{\varepsilon}(E\cap U_{\delta}(y_{2},-\xi))}\setminus\{x\}\neq\varnothing

(see Proposition 3.1 for clarification). In both cases there exists some p(x)<0p(x)<0 for which

(5.6) Q(x):={x+sξ+t(xy1):p(x)<s<0,ε/2<t<ε/2}intEε.Q(x):=\big{\{}x+s\xi+t(x-y_{1})\,:\,p(x)<s<0,\,-\varepsilon/2<t<\varepsilon/2\big{\}}\subset\mathrm{int}\,E_{\varepsilon}.

One can then define Ax:=Up(x)/3(x,ξ)A_{x}:=U_{p(x)/3}(x,-\xi). Note that for all xCx\in C the set AxA_{x} is open and has a positive surface area 2(Ax)>0\mathcal{H}^{2}(A_{x})>0. Furthermore, AxQ(x)intEεA_{x}\subset Q(x)\subset\textrm{int}\,E_{\varepsilon} and it follows from the construction that if zxz\neq x is any other one-sided chain singularity, its distance from xx satisfies |zx|p(x)|z-x|\geq p(x), implying AxAz=A_{x}\cap A_{z}=\varnothing. The sets AxA_{x} are thus pair-wise disjoint, open and contained in some bounded ball BR(0)B_{R}(0) due to the compactness of EεE_{\varepsilon}. Hence the sum xC2(Ax)\sum_{x\in C}\mathcal{H}^{2}(A_{x}) of their surface areas is finite, from which the result follows by the counting argument discussed in the lead-up to Lemma 5.4. ∎

5.1.1. Proof of Theorem 2

We conclude this section with the proof of Theorem 2, which combines Lemmas 5.25.4 and 5.6 into one statement.

Theorem 2 (Countable sets of singularities).

For a compact set E2E\subset\mathbb{R}^{2}, the number of wedges (S1), sharp singularities (S2, S3 and S8) and one-sided shallow singularities (S4) and chain singularities (S6) on Eε\partial E_{\varepsilon} is at most countably infinite.

Proof.

Consider the collection {Ui}iI\{U_{i}\}_{i\in I} of the connected components of EεcE_{\varepsilon}^{c}. Since EE is assumed to be compact, EBR(0)E\subset B_{R}(0) for some R>0R>0. It follows that all but one (denote this by UjU_{j}) of the connected components UiU_{i} are bounded, so that

iI{j}UiBR(0).\bigcup_{i\in I\setminus\{j\}}U_{i}\subset B_{R}(0).

Following the counting argument presented immediately before the statement of Lemma 5.4, this implies that the index set II is at most countably infinite. By definition every sharp singularity (S2, S3, S8) xEεx\in\partial E_{\varepsilon} satisfies xiIUix\in\bigcup_{i\in I}\partial U_{i}. It follows then from Lemma 5.3 that the set of sharp singularities on Eε\partial E_{\varepsilon} is countable as a countable union of finite sets. Finally, Lemmas 5.2, 5.4 and 5.6 guarantee that the number of wedges (S1) and one-sided shallow (S4) and chain (S6) singularities (respectively) are at most countably infinite, and the proof is complete. ∎

5.2. Chain Singularities Form a Totally Disconnected Set

We conclude the paper by showing that the set 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) of chain singularities (types S6–S8) is closed and totally disconnected. This implies that 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) is nowhere dense, meaning that it is small in the topological sense, even though it may have a positive one-dimensional Hausdorff measure on the boundary.

Before presenting the proof of the above result, we provide a concrete example of a set E2E\subset\mathbb{R}^{2} and ε>0\varepsilon>0 for which the one-dimensional Hausdorff measure of the set of chain-chain singularities on Eε\partial E_{\varepsilon} is positive. Essentially, we analyse [27, Example 2.2] from the geometric point of view.

Example 5.7 (A set of chain singularities with positive measure).

Let C[0,1]C\subset[0,1] be a 'fat' Cantor set (a Cantor set with positive one-dimensional Hausdorff measure) and consider the set E:={(s,t)2:sC,t{0,1}}E:=\big{\{}(s,t)\in\mathbb{R}^{2}\,:\,s\in C,\,t\in\{0,1\}\big{\}}. The Cantor set is obtained by removing from the interval [0,1][0,1] a certain countable collection :={In:n}\mathcal{I}:=\{I_{n}\,:\,n\in\mathbb{N}\} of open subintervals. By construction CC is totally disconnected and thus contains no intervals itself, and since it is uncountable, most of the points sCs\in C do not lie on the boundary of any of the removed intervals. Denote the collection of these points by C:=CnInC^{*}:=C\setminus\bigcup_{n\in\mathbb{N}}\partial I_{n}. By construction, every sCs\in C^{*} is however an accumulation point of nIn\bigcup_{n\in\mathbb{N}}I_{n}. Since the sets InI_{n} are open, it is clear that for ε=1/2\varepsilon=1/2 the sets Vn:={(s,1/2):sIn}V_{n}:=\{(s,1/2)\,:\,s\in I_{n}\} satisfy VnEεcV_{n}\subset E_{\varepsilon}^{c} for all nn\in\mathbb{N}. Thus, for every point xA:={(s,1/2):sC}Eεx\in A:=\{(s,1/2)\,:\,s\in C^{*}\}\subset\partial E_{\varepsilon} there exists a sequence (wm)m=1Eεc(w_{m})_{m=1}^{\infty}\subset E_{\varepsilon}^{c}, where wm=(sm,1/2)Vmw_{m}=(s_{m},1/2)\in V_{m} for some smIn(m)s_{m}\in I_{n(m)}, and smss_{m}\to s as mm\to\infty. We can also assume that In(m)In(m)I_{n(m)}\neq I_{n(m^{\prime})} whenever mmm\neq m^{\prime}. For the connected components WmVmW_{m}\supset V_{m} of EεcE_{\varepsilon}^{c} this implies WmWmW_{m}\neq W_{m^{\prime}} whenever mmm\neq m^{\prime}. It is easy to see that condition (ii) in Proposition 4.5 thus holds true for all xAx\in A so that A𝒞(Eε)A\subset\mathcal{C}(\partial E_{\varepsilon}). Hence 1(𝒞(Eε))1(A)=1(C)>0\mathcal{H}^{1}(\mathcal{C}(\partial E_{\varepsilon}))\geq\mathcal{H}^{1}(A)=\mathcal{H}^{1}(C^{*})>0 due to the translation invariance of Hausdorff measure.

5.2.1. Proof of Theorem 3

The proof of our third main result builds on many of the results presented in the previous sections. To show that the set of chain singularities is closed, we combine Proposition 3.9 regarding the connectedness of the complement EεcE_{\varepsilon}^{c} near wedges and xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E) with the characterisation of chain singularities provided by Proposition 4.5. The second part of the proof also makes use of the basic results established in Section 2.2.

Theorem 3 (The set of chain singularities is closed and totally disconnected).

For any compact set E2E\subset\mathbb{R}^{2} and ε>0\varepsilon>0, the set 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) of chain singularities is closed and totally disconnected.

Proof.

We begin by showing that the complement Eε𝒞(Eε)\partial E_{\varepsilon}\setminus\mathcal{C}(\partial E_{\varepsilon}) is open. To this end, consider some xEε𝒞(Eε)x\in\partial E_{\varepsilon}\setminus\mathcal{C}(\partial E_{\varepsilon}). If xx is a wedge (S1) or if xUnpε(E)x\in\mathrm{Unp_{\varepsilon}}(E), Proposition 3.9 implies that there exists some neighbourhood Br(x)B_{r}(x) and a connected subset VxEεcV_{x}\subset E_{\varepsilon}^{c} for which

(5.7) Br(x)Eεc=Br(x)Vx.B_{r}(x)\cap E_{\varepsilon}^{c}=B_{r}(x)\cap V_{x}.

One the other hand, Proposition 4.5 states that each chain singularity x𝒞(Eε)x\in\mathcal{C}(\partial E_{\varepsilon}) is associated with a sequence (Vn)n=1Eεc(V_{n})_{n=1}^{\infty}\subset E_{\varepsilon}^{c} of disjoint connected components of the complement EεcE_{\varepsilon}^{c}, for which distH(x,Vn)0\mathrm{dist}_{H}(x,V_{n})\to 0 as nn\to\infty. Equation (5.7) hence implies that Br(x)𝒞(Eε)=B_{r}(x)\cap\mathcal{C}(\partial E_{\varepsilon})=\varnothing. Similarly, for a sharp singularity (type S2) or a sharp-sharp singularity (type S3), Proposition 4.2 implies the existence of a neighbourhood Br(x)B_{r}(x) for which Br(x)𝒞(Eε)=B_{r}(x)\cap\mathcal{C}(\partial E_{\varepsilon})=\varnothing. Hence Eε𝒞(Eε)\partial E_{\varepsilon}\setminus\mathcal{C}(\partial E_{\varepsilon}) is open on the boundary.

To demonstrate that 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) is totally disconnected, we show that for any two chain singularities x,z𝒞(Eε)x,z\in\mathcal{C}(\partial E_{\varepsilon}) there exist disjoint open sets Ax,Az2A_{x},A_{z}\subset\mathbb{R}^{2} for which xAx,zAzx\in A_{x},z\in A_{z} and 𝒞(Eε)(AxAz)Eε\mathcal{C}(\partial E_{\varepsilon})\subset(A_{x}\cup A_{z})\cap\partial E_{\varepsilon}. More specifically, we will consider for each x𝒞(Eε)x\in\mathcal{C}(\partial E_{\varepsilon}) and s10s2s_{1}\leq 0\leq s_{2} the sets

(5.8) Ax,ξ(s1,s2):={x+sξ+t(xy1):s1<s<s2,ε/2<t<ε/2}A_{x,\xi}(s_{1},s_{2}):=\big{\{}x+s\xi+t(x-y_{1})\,:\,s_{1}<s<s_{2},\,\,-\varepsilon/2<t<\varepsilon/2\big{\}}

and show that for each z𝒞(Eε){x}z\in\mathcal{C}(\partial E_{\varepsilon})\setminus\{x\} there exist ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}) and s1<0<s2s_{1}<0<s_{2} for which Ax,ξ(s1,s2)𝒞(Eε)=\partial A_{x,\xi}(s_{1},s_{2})\cap\mathcal{C}(\partial E_{\varepsilon})=\varnothing and zAx,ξ(s1,s2)z\notin A_{x,\xi}(s_{1},s_{2}).

Given that xx is a chain singularity, Proposition 4.2 implies that for r>0r>0 the boundary region EεUr(x,ξ)\partial E_{\varepsilon}\cap U_{r}(x,\xi) exhibits 'chain-type' geometry near xx for at least one extremal outward direction ξΞxext(Eε)\xi\in\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon}). We begin by assuming that ξ\xi is such a direction, and consider the corresponding sets Ax,ξ(0,s)A_{x,\xi}(0,s) for s>0s>0. Writing R(z):=zxR(z):=\left\lVert z-x\right\rVert, our aim is to find some s<R(z)/2s<R(z)/2 for which Ax,ξ(0,s)𝒞(Eε)={x}\partial A_{x,\xi}(0,s)\cap\mathcal{C}(\partial E_{\varepsilon})=\{x\}. To this end, let 𝒢(x)\mathcal{G}(x) be a local boundary representation at xx and let fξ,y1,fξ,y2:[0,ε/2]f^{\xi,y_{1}},f^{\xi,y_{2}}:[0,\varepsilon/2]\to\mathbb{R} be the continuous functions for which

gξ,yi(s)=x+sξ+fξ,yi(s)(xyi)g_{\xi,y_{i}}(s)=x+s\xi+f^{\xi,y_{i}}(s)(x-y_{i})

for every gξ,yi𝒢(x)g_{\xi,y_{i}}\in\mathcal{G}(x), i{1,2}i\in\{1,2\} and all s[0,ε/2]s\in[0,\varepsilon/2]. As argued in the proof of Proposition 4.2 (ii), there exist sequences (sn)n=1+(s_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} and (pn)n=1+(p_{n})_{n=1}^{\infty}\subset\mathbb{R}_{+} for which

  • pn<snpn1<sn1p_{n}<s_{n}\leq p_{n-1}<s_{n-1} for all nn\in\mathbb{N},

  • fξ,y1(s)+fξ,y2(s)=0f^{\xi,y_{1}}(s)+f^{\xi,y_{2}}(s)=0 for all s(sn)n=1(pn)n=1s\in(s_{n})_{n=1}^{\infty}\cup(p_{n})_{n=1}^{\infty}, and

  • fξ,y1(s)+fξ,y2(s)<0f^{\xi,y_{1}}(s)+f^{\xi,y_{2}}(s)<0 for all s(pn,sn)s\in(p_{n},s_{n}) and nn\in\mathbb{N}.

It follows that the open set

Vn:={τgξ,y1(s)+(1τ)gξ,y2(s):τ(0,1),s(pn,sn)}V_{n}:=\big{\{}\tau g_{\xi,y_{1}}(s)+(1-\tau)g_{\xi,y_{2}}(s)\,:\,\tau\in(0,1),\,s\in(p_{n},s_{n})\big{\}}

is a connected component of the complement EεcE_{\varepsilon}^{c} for all nn\in\mathbb{N}. Consequently there exists some NN\in\mathbb{N} for which distH(Vn,x)<R(z)/2\mathrm{dist}_{H}(V_{n},x)<R(z)/2 whenever nNn\geq N. The definition of the sets VnV_{n} implies that for each nNn\geq N, the boundary point xn(1):=gξ,y1(pn+sn2)VnEεx_{n}^{(1)}:=g_{\xi,y_{1}}\big{(}\frac{p_{n}+s_{n}}{2}\big{)}\in\partial V_{n}\subset\partial E_{\varepsilon} has an outward direction aligned with the vector

ηn(1):=gξ,y2(pn+sn2)gξ,y1(pn+sn2)=(fξ,y1(pn+sn2)+fξ,y2(pn+sn2))(xy1)\eta_{n}^{(1)}:=g_{\xi,y_{2}}\left(\frac{p_{n}+s_{n}}{2}\right)-g_{\xi,y_{1}}\left(\frac{p_{n}+s_{n}}{2}\right)=-\left(f^{\xi,y_{1}}\left(\frac{p_{n}+s_{n}}{2}\right)+f^{\xi,y_{2}}\left(\frac{p_{n}+s_{n}}{2}\right)\right)(x-y_{1})

(the reader is encouraged to compare this with equation (4.3) in the proof of Proposition 4.2), and the analogous expression (obtained by replacing the roles of y1y_{1} and y2y_{2}) holds true for xn(2)x_{n}^{(2)} and ηn(2)\eta_{n}^{(2)} defined similarly. We claim that there exists some nNn\geq N, for which xn(i)𝒞(Eε)x_{n}^{(i)}\notin\mathcal{C}(\partial E_{\varepsilon}) for i{1,2}i\in\{1,2\}. Assume this were not the case. Then it follows from the definition of chain-singularities and Proposition 2.12 that for at least one i{1,2}i\in\{1,2\} we have ηn(i)Ξxn(i)ext(Eε)\eta_{n}^{(i)}\in\Xi_{x_{n}^{(i)}}^{\mathrm{ext}}(E_{\varepsilon}) for infinitely many nNn\geq N. By virtue of the definition of the sets VnV_{n} as regions between the graphs gξ,yig_{\xi,y_{i}}, and since xn(i)VnEεx_{n}^{(i)}\in\partial V_{n}\subset\partial E_{\varepsilon} for all nn\in\mathbb{N}, it follows from Proposition 2.14 that ξ=limn(xn(i)x)/xn(i)x\xi=\lim_{n\to\infty}\big{(}x_{n}^{(i)}-x\big{)}\big{/}\big{\|}x_{n}^{(i)}-x\big{\|} for i{1,2}i\in\{1,2\}. But then, due to Lemma 2.13 (ii)(b), we have

0=xyiε,ξ=limnηn(i),ξ=1,0=\left\langle\frac{x-y_{i}}{\varepsilon},\,\xi\right\rangle=\lim_{n\to\infty}\big{\langle}\eta_{n}^{(i)},\xi\big{\rangle}=1,

which is impossible. Thus, there exists some nn\in\mathbb{N} for which xn(i)𝒞(Eε)x_{n}^{(i)}\notin\mathcal{C}(\partial E_{\varepsilon}) for i{1,2}i\in\{1,2\}, which in turn implies that Ax,ξ(0,(pn+sn)/2)𝒞(Eε)={x}\partial A_{x,\xi}\big{(}0,(p_{n}+s_{n})/2\big{)}\cap\mathcal{C}(\partial E_{\varepsilon})=\{x\}, since

Ax,ξ(0,pn+sn2)Eε={x,xn(1),xn(2)}.\partial A_{x,\xi}\left(0,\frac{p_{n}+s_{n}}{2}\right)\cap\partial E_{\varepsilon}=\big{\{}x,x_{n}^{(1)},x_{n}^{(2)}\big{\}}.

In the remainder of the proof we consider one by one the cases of one-sided chain (S6), chain-chain (S7) and sharp-chain (S8) singularities, and identify the sets AxA_{x} and AzA_{z} mentioned in the beginning of the proof.

(i) Assume xx is a one-sided chain singularity (S6), so that Ξxext(Eε)={ξ}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi\} for some ξS1\xi\in S^{1}. By the argument presented above, there exists some 0<s2<R(z)/20<s_{2}<R(z)/2 for which Ax,ξ(0,s2)𝒞(Eε)={x}\partial A_{x,\xi}\big{(}0,s_{2}\big{)}\cap\mathcal{C}(\partial E_{\varepsilon})=\{x\}. On the other hand, according to the reasoning presented in the proof of Lemma 5.6, the set

Qp(x):={x+sξ+t(xy1):p<s<0,ε/2<t<ε/2}Q_{p}(x):=\big{\{}x+s\xi+t(x-y_{1})\,:\,p<s<0,\,-\varepsilon/2<t<\varepsilon/2\big{\}}

satisfies Qp(x)intEεQ_{p}(x)\subset\mathrm{int}\,E_{\varepsilon} for some p<0p<0 (see equation (5.6)). By setting s1:=min{p,R(z)/2}s_{1}:=-\mathrm{min}\big{\{}p,R(z)/2\big{\}} it follows then that Ax,ξ(s1,s2)𝒞(Eε)=\partial A_{x,\xi}(s_{1},s_{2})\cap\mathcal{C}(\partial E_{\varepsilon})=\varnothing and we may define Ax:=Ax,ξ(s1,s2)A_{x}:=A_{x,\xi}(s_{1},s_{2}) and Az:=(2Ax¯)A_{z}:=\big{(}\mathbb{R}^{2}\setminus\overline{A_{x}}\big{)}.

(ii) Assume then that xx is a chain-chain singularity (S7) with Ξxext(Eε)={ξ,ξ}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi,-\xi\} for some ξS1\xi\in S^{1}. Since now both of the extremal outward directions ξ\xi and ξ-\xi are associated with 'chain'-type geometry, one can again utilise the argument above in order to choose for ξ1:=ξ\xi_{1}:=\xi and ξ2:=ξ\xi_{2}:=-\xi the corresponding s1,s2>0s_{1},s_{2}>0 for which Ax,ξi(0,si)𝒞(Eε)={x}\partial A_{x,\xi_{i}}\big{(}0,s_{i}\big{)}\cap\mathcal{C}(\partial E_{\varepsilon})=\{x\} and si<R(z)/2s_{i}<R(z)/2 for i{1,2}i\in\{1,2\}. By setting s3=min{s1,s2}s_{3}=\mathrm{min}\{s_{1},s_{2}\} we may define Ax:=Ax,ξ(s3,s3)A_{x}:=A_{x,\xi}(-s_{3},s_{3}) and Az:=(2Ax¯)A_{z}:=\big{(}\mathbb{R}^{2}\setminus\overline{A_{x}}\big{)}.

(iii) Finally, assume that xx is a sharp-chain singularity (S8) with Ξxext(Eε)={ξ,ξ}\Xi_{x}^{\mathrm{ext}}(E_{\varepsilon})=\{\xi,-\xi\} for some ξS1\xi\in S^{1}. We may assume that ξ\xi is associated with 'chain'-type geometry, so that once again we have Ax,ξ(0,s2)𝒞(Eε)={x}\partial A_{x,\xi}\big{(}0,s_{2}\big{)}\cap\mathcal{C}(\partial E_{\varepsilon})=\{x\} for some 0<s2<R(z)/20<s_{2}<R(z)/2 due to the arguments presented above. For the direction ξ-\xi, Proposition 4.2 (i) implies that there exists some r>0r>0 for which

EεcUr(x,ξ)=VUr(x,ξ)=0<s<rx+s(α(s),β(s))S1,E_{\varepsilon}^{c}\cap U_{r}(x,-\xi)=V\cap U_{r}(x,-\xi)=\bigcup_{0<s<r}x+s\big{(}\alpha(s),\beta(s)\big{)}_{S^{1}},

where VV is the unique connected component of EεcE_{\varepsilon}^{c} intersecting Ur(x,ξ)U_{r}(x,-\xi), α(s),β(s)S1\alpha(s),\beta(s)\in S^{1} for all s(0,r)s\in(0,r) and α(s),β(s)ξ\alpha(s),\beta(s)\to-\xi as s0s\to 0. It follows then from the definition of chain singularity that Ur(x,ξ)𝒞(Eε)=U_{r}(x,-\xi)\cap\mathcal{C}(\partial E_{\varepsilon})=\varnothing. By setting s1=min{r/2,ε/2,R(z)/2}s_{1}=-\mathrm{min}\{r/2,\varepsilon/2,R(z)/2\} we may thus define Ax:=Ax,ξ(s1,s2)A_{x}:=A_{x,\xi}(s_{1},s_{2}) and Az:=(2Ax¯)A_{z}:=\big{(}\mathbb{R}^{2}\setminus\overline{A_{x}}\big{)} and the proof is complete. ∎

The set \mathcal{I} of inaccessible singularities (types S6 and S7) inherits the properties of being totally disconnected and nowhere dense, but it may generally fail to be closed. Since the set 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) on the other hand is compact and separable as a subset of 2\mathbb{R}^{2}, it follows from the Cantor-Bendixson Theorem (see [19, Thm. 6.4]) that whenever the cardinality of the chain-chain singularities (type S7) is uncountable, the set 𝒞(Eε)\mathcal{C}(\partial E_{\varepsilon}) can be written as a disjoint union 𝒞(Eε)=CP\mathcal{C}(\partial E_{\varepsilon})=C\cup P, where CC is homeomorphic to the Cantor set and PP is countable. For further information on totally disconnected spaces, see [19].

Acknowledgements

This project has received funding from the European Union's Horizon 2020 research and innovation Programme under the Marie Skłodowska-Curie grant agreement no. 643073. The authors express their gratitude to Gabriel Fuhrman, Vadim Kulikov, Tuomas Orponen, Sebastian van Strien and Dmitry Turaev for many useful discussions and valuable comments regarding earlier versions of this article.

References

  • [1] L. Ambrosio, N. Fusco, and D. Pallara, Functions of Bounded Variation and Free Discontinuity Problems, Oxford University Press, 2000.
  • [2] J.-P. Aubin and H. Frankowska, Set-Valued Analysis, Birkhäuser, Boston, 1990.
  • [3] F. Bigolin and G. H. Greco, Geometric characterizations of C1{C}^{1}-manifolds in euclidean spaces by tangent cones, J. Math. Anal. Appl. 396 (2012), 145–163.
  • [4] A. Blokh, M. Misiurewicz, and L. Oversteegen, Sets of constant distance from a compact set in 2-manifolds with a geodesic metric, Proc. Amer. Math. Soc. 137 (2009), no. 2, 733–743.
  • [5] G. Bouligand, Sur les surfaces dépourvues de points hyperlimites, Ann. Soc. Polon. Math. 9 (1930), 32–41.
  • [6] by same author, Introduction á la géométrie infinitésimale directe, Gauthier-Villars, 1932.
  • [7] M. Brown, Sets of constant distance from a planar set, Michigan Math. J. 19 (1972), 321–323.
  • [8] F. H. Clarke, Generalized gradients and applications, Trans. Am. Math. Soc. 205 (1975), 247–262.
  • [9] F. Colonius and W. Kliemann, The Dynamics of Control, Birkhäuser, Boston, 2000.
  • [10] M. D. Donsker and S. R. S. Varadhan, Asymptotics for the Wiener Sausage, Comm. Pure and Appl. Math. 28 (1975), 525–565.
  • [11] P. Erdős, Some remarks on the measurability of certain sets, Bull. Amer. Math. Soc. 51 (1945), no. 10, 728–731.
  • [12] K. J. Falconer, The geometry of fractal sets, Cambridge University Press, 1985.
  • [13] H. Federer, Curvature measures, Trans. Amer. Math. Soc. 93 (1959), 418–491.
  • [14] S. Ferry, When ε\varepsilon-boundaries are manifolds, Fund. Math. 90 (1976), 199–210.
  • [15] J.H.G. Fu, Tubular neighborhoods in Euclidian spaces, Duke Mathematical Journal 52 (1985), no. 4, 1025–1046.
  • [16] Y. Hamana, On the expected volume of the Wiener sausage, J. Math. Soc. Japan 62 (2010), no. 1, 1113–1136.
  • [17] A. Kac and J. M. Luttinger, Bose-Einstein condensation in the presence of impurities, J. Math. Phys. 14 (1973), 1626–1628.
  • [18] A. Käenmäki, J. Lehrbäck, and M. Vuorinen, Dimensions, Whitney covers, and tubular neighborhoods, Indiana Univ. Math. J. 62 (2013), no. 6, 1861–1889.
  • [19] A.S. Kechris, Classical descriptive set theory, Graduate Texts in Mathematics, pp. 31–32, 35–36, Springer-Verlag, 1995.
  • [20] B. Klartag, Eldan's Stochastic Localization and Tubular Neighborhoods of Complex-Analytic Sets, The Journal of Geometric Analysis 28 (2018), 2008–2018.
  • [21] J.S.W. Lamb, M. Rasmussen, and C. Rodrigues, Topological bifurcations of minimal invariant sets for set-valued dynamical systems, Proceedings of the American Mathematical Society 143 (2015), no. 9, 3927–3937.
  • [22] S. Nekovar and G. Pruessner, A Field-Theoretic Approach to the Wiener Sausage, Journal of Statistical Physics 163 (2016), 604–641.
  • [23] I. Ya. Oleksiv and N. I. Pesin, Finiteness of the Hausdorff measure of level sets of bounded subsets in Euclidian space, Mat. Zametki 37 (1985), 422–431, (in Russian).
  • [24] A. Przeworski, An upper bound on density for packings of collars about hyperplanes in n\mathbb{H}^{n}, Geom. Dedicata 163 (2013), 193–213.
  • [25] J. Rataj, E. Spodarev, and D. Meschenmoser, Approximations of the Wiener Sausage and Its Curvature Measures, The Annals of Applied Probability 19 (2009), no. 5, 1840–1859.
  • [26] J. Rataj and E. Spodarev V. Schmidt, On the Expected Surface Area of the Wiener Sausage, Mathematische Nachrichten 282 (2009), no. 4, 591–603.
  • [27] J. Rataj and S. Winter, On volume and surface area of parallel sets, Indiana University Mathematics Journal 59 (2010), no. 5, 1661–1685.
  • [28] J. Rataj and L. Zajíček, Properties of distance functions on convex surfaces and applications, Czech. Math. J. 61 (2011), 247–269.
  • [29] by same author, Smallness of the critical values of distance functions in two-dimensional euclidean and riemannian spaces, Mathematika 66 (2020), 297–324.
  • [30] R. T. Rockafellar, Clarke's tangent cones and the boundaries of closed sets in n\mathbb{R}^{n}, Nonlinear Analysis: Theory, Methods and Applications 3 (1979), no. 1, 145–154.
  • [31] R. T. Rockafellar and R. J.-B. Wets, Variational analysis, Springer-Verlag, 1998.
  • [32] F. Spitzer, Electrostatic Capacity, Heat Flow, and Brownian Motion, Z. Wahrscheinlichkeitstheorie 3 (1964), 110–121.
  • [33] L.L. Stachó, On the volume function of parallel sets, Acta Sci. Math. 38 (1976), 365–374.