• Non ci sono risultati.

Bounding Regions to Plane Steepest Descent Curves of Quasi Convex families

N/A
N/A
Protected

Academic year: 2021

Condividi "Bounding Regions to Plane Steepest Descent Curves of Quasi Convex families"

Copied!
18
0
0

Testo completo

(1)

Research Article

Bounding Regions to Plane Steepest Descent Curves of

Quasiconvex Families

Marco Longinetti,

1

Paolo Manselli,

1

and Adriana Venturi

2

1Dipartimento di Matematica e Informatica Ulisse Dini, Universit`a degli Studi di Firenze, Viale Morgagni 67/a, 50134 Firenze, Italy 2Dipartimento GESAAF, Universit`a degli Studi di Firenze, Piazzale delle Cascine 15, 50144 Firenze, Italy

Correspondence should be addressed to Marco Longinetti; marco.longinetti@unifi.it Received 26 February 2016; Accepted 17 April 2016

Academic Editor: Wenyu Sun

Copyright © 2016 Marco Longinetti et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Two-dimensional steepest descent curves (SDC) for a quasiconvex family are considered; the problem of their extensions (with

constraints) outside of a convex body𝐾 is studied. It is shown that possible extensions are constrained to lie inside of suitable

bounding regions depending on𝐾. These regions are bounded by arcs of involutes of 𝜕𝐾 and satisfy many inclusions properties.

The involutes of the boundary of an arbitrary plane convex body are defined and written by their support function. Extensions SDC of minimal length are constructed. Self-contracting sets (with opposite orientation) are considered: necessary and/or sufficient conditions for them to be subsets of SDC are proved.

1. Introduction

Let𝑢 be a smooth function defined in a convex body Ω ⊂ R𝑛.

Let𝐷𝑢(𝑥) ̸= 0 in {𝑥 ∈ Ω : 𝑢(𝑥) > min 𝑢}. A classical steepest

descent curve of𝑢 is a rectifiable curve 𝑠 → 𝑥(𝑠) solution to

𝑑𝑥

𝑑𝑠 = −

𝐷𝑢

|𝐷𝑢|(𝑥 (𝑠)) . (1)

Classical steepest descent curves are the integral curves of a unit field normal to the sublevel sets of the given smooth

function𝑢. We are interested in “generalized” steepest descent

curves that are integral curves to a unit field normal to a

nested family of convex sets {Ω𝑡} (see Definition 5); {Ω𝑡}

will be called a quasiconvex family as in [1]. Sharp bounds about the length of the steepest descent curves for a quasi convex family have been proved in [2–4]. The geometry of these curves, equivalent definitions, related questions and generalizations have been studied in [5–9].

In the present work generalized steepest descent curves for

a quasiconvex family (SDC for short) are defined as bounded

oriented rectifiable curves𝛾 ⊂ R𝑛, with a locally Lipschitz

continuous parameterization 𝑇 ∋ 𝑡 → 𝑥(𝑡), with ascent

parameter, satisfying

⟨ ̇𝑥 (𝑡) , 𝑥 (𝜏) − 𝑥 (𝑡)⟩ ≤ 0, a.e. 𝑡 ∈ 𝑇, ∀𝜏 ≤ 𝑡; (2)

⟨⋅, ⋅⟩ is the scalar product in R𝑛. Let ordering⪯ be chosen on

𝛾, according to the orientation; let us denote

𝛾𝑥= {𝑦 ∈ 𝛾 : 𝑦 ⪯ 𝑥} . (3)

In [8, Theorem 4.10], the SDC are characterized in an equiv-alent way as self-distancing curves, namely, oriented (⪯)

continuous curves with the property that the distance of𝑥

to an arbitrarily fixed previous point𝑥1is not decreasing:

𝑥1⪯ 𝑥2⪯ 𝑥3󳨐⇒

󵄨󵄨󵄨󵄨𝑥2− 𝑥1󵄨󵄨󵄨󵄨 ≤ 󵄨󵄨󵄨󵄨𝑥3− 𝑥1󵄨󵄨󵄨󵄨 ∀𝑥1, 𝑥2, 𝑥3∈ 𝛾.

(4) Thus steepest descent curves are self-distancing curves and both denoted SDC. In [8] self-distancing curves are called self-expanding curves. With the opposite orientation these curves have been also introduced, studied, and called self-approaching curves (see [2]) or self-contracting curves (see [7]).

Volume 2016, Article ID 4873276, 17 pages http://dx.doi.org/10.1155/2016/4873276

(2)

In our work we are interested in the behaviour and

properties of a plane SDC𝛾 beyond its final point 𝑥0. One of

the principal goals of the paper is to show that conditions (2)

and (4) imply constraints for possible extensions of curve𝛾

beyond𝑥0; these constraints are written as bounding regions

for the possible extensions of𝛾𝑥0.

An important property that will be used later is the

property of distancing from a set𝐴.

Definition 1. Given a set𝐴, an absolutely continuous curve 𝛾,

𝑇 ∋ 𝑡 → 𝑥(𝑡) has the distancing from 𝐴 property if it satisfies

⟨ ̇𝑥 (𝑡) , 𝑦 − 𝑥 (𝑡)⟩ ≤ 0, a.e. 𝑡 ∈ 𝑇, ∀𝑦 ∈ 𝐴 . (5)

Let us outline the content of our work. In Section 2 introductory definitions are given and covering maps for the boundary of a plane convex set, needed for later use, are introduced. In Section 3 the involutes of the boundary of a plane convex body are introduced and some of their properties are proved.

In Section 4 plane regions depending on the convex hull

of𝛾𝑥0have been defined; these regions fence in or fence out

the possible extensions of 𝛾𝑥0. The boundary of these sets

consists of arcs of involutes of convex bodies, constructed in Section 3. As an application, in Section 4.1 the following

problem has been studied: given a convex set𝐾, 𝑥0∈ 𝜕𝐾, 𝑥1

𝐾, is it possible to construct SDC joining 𝑥0to𝑥1, satisfying

the distancing from K property? Minimal properties of this construction have been introduced and studied. In Section 5

sets of points more general than SDC are studied. A set𝜎 ⊂

R2(not necessarily a curve) of ordered points satisfying (4)

will be called self-distancing set; see also Definition 2; with

the opposite order,𝜎 was called in [6] self-contracting set

and many properties of these sets, as only subsets of self-contracting curves, were there obtained. Another goal of the paper is the solution to the following question: given a

self-distancing set𝜎 ⊂ R2 does a steepest descent curve𝛾 ⊃

𝜎 exist? In Section 5 examples, necessary and/or sufficient

conditions are given when𝜎 consists of a finite or countable

number of points 𝑥𝑖 ∈ R2 and/or steepest descent curves

𝛾𝑖⊂ R2.

In the present work the two-dimensional case is studied.

Similar results for the 𝑛-dimensional case are an open

problem stated at the end of the work.

2. Preliminaries and Definitions

Let

𝐵 (𝑧, 𝜌) = {𝑥 ∈ R𝑛: |𝑥 − 𝑧| < 𝜌} ,

𝑆𝑛−1= 𝜕𝐵 (0, 1)

𝑛 ≥ 2.

(6)

A nonempty, compact convex set𝐾 of R𝑛 will be called a

convex body. From now on,𝐾 will always be a convex body not reduced to a point. Int(𝐾) and 𝜕𝐾 denote the interior of 𝐾 and

the boundary of𝐾, |𝜕𝐾| denotes its length, cl(𝐾) is the closure

of𝐾, Aff(𝐾) will be the smallest affine space containing 𝐾,

and relint𝐾 and 𝜕rel𝐾 are the corresponding subsets in the

topology of Aff(𝐾). For every set 𝑆 ⊂ R𝑛, co(𝑆) is the convex

hull of𝑆.

Let𝑞 ∈ 𝐾; the normal cone at 𝑞 to 𝐾 is the closed convex

cone:

𝑁𝐾(𝑞) = {𝑥 ∈ R𝑛: ⟨𝑥, 𝑦 − 𝑞⟩ ≤ 0 ∀𝑦 ∈ 𝐾} . (7)

When𝑞 ∈ Int(𝐾), then 𝑁𝐾(𝑞) reduces to zero.

The tangent cone or support cone of K at a point𝑞 ∈ 𝜕𝐾

is given by

𝑇𝐾(𝑞) = cl ( ⋃

𝑦∈𝐾

{𝑠 (𝑦 − 𝑞) : 𝑠 ≥ 0}) . (8)

In two dimensions cones will be called sectors.

Let𝐾 be a convex body and let 𝑝 be a point. A simple cap

body𝐾𝑝is

𝐾𝑝= ⋃

0≤𝜆≤1{𝜆𝐾 + (1 − 𝜆) 𝑝} = co (𝐾 ∪ {𝑝}) . (9)

Cap bodies properties can be found in [10, 11].

2.1. Self-Distancing Sets and Steepest Descent Curves. Let us

recall the following definitions.

Definition 2. Let us call self-distancing set a bounded subset𝜎

ofR𝑛, linearly ordered (by⪯), with the property

𝑥1⪯ 𝑥2⪯ 𝑥3󳨐⇒

󵄨󵄨󵄨󵄨𝑥2− 𝑥1󵄨󵄨󵄨󵄨 ≤ 󵄨󵄨󵄨󵄨𝑥3− 𝑥1󵄨󵄨󵄨󵄨, 𝑥1, 𝑥2, 𝑥3∈ 𝜎.

(10) The self-distancing sets have been introduced in [6] with

the opposite order. If a self-distancing set 𝜎 is a closed

connected set, not reduced to a point, then it can be proved

that 𝜎 is the support of a steepest descent curve 𝛾 (see [8,

Theorem 4.10, Theorem 4.8]) and it will also be called a

self-distancing curve𝛾.

The short name SDC will be used both for self-distancing curves and for steepest descent curves in all the paper.

Definition 3. Let𝐾 be a convex body; 𝛾 ⊂ R2\ relint 𝐾 will

be called a self-distancing curve from𝐾 (denoted SDC𝐾) if

(i)𝛾 is a self-distancing curve,

(ii)𝛾 ∩ 𝜕rel𝐾 ̸= 0,

(iii)𝛾 has the property

𝑥 ⪯ 𝑥1󳨐⇒

󵄨󵄨󵄨󵄨𝑥 − 𝑦󵄨󵄨󵄨󵄨 ≤ 󵄨󵄨󵄨󵄨𝑥1− 𝑦󵄨󵄨󵄨󵄨 ∀𝑦 ∈ 𝐾, ∀𝑥, 𝑥1∈ 𝛾.

(11)

When (ii) does not hold, that is𝛾 ∩ 𝜕rel𝐾 = 0, 𝛾 will be called

a deleted self-distancing curve from𝐾.

Remark 4. Let𝛾 be SDC𝐾, since𝛾 has an absolutely continu-ous parameterization ([8, Theorem 4.10, Theorem 4.8]), thus

(3)

If𝛾 is SDC and 𝑥 ∈ 𝛾 then 𝛾 \ 𝛾𝑥 is SDCco(𝛾𝑥). That is,

the tangent vector ̇𝑥(𝑡) to 𝛾 is in the normal cone at 𝑥(𝑡) to

the related convex setΩ𝑡 fl co(𝛾𝑥(𝑡)). This condition is a

generalization of the classicals steepest descent curves that are integral curves to a unit field normal to smooth quasiconvex families.

Nested families of convex sets have been introduced and studied by de Finetti [12] and Fenchel [1]. Let us recall some definitions.

Definition 5. Let𝑇 be a real interval. A convex stratification

(see [12]) is a nonempty familyK of convex bodies Ω𝑡⊂ R𝑛,

𝑡 ∈ 𝑇 ⊂ R, linearly strictly ordered by inclusion (Ω1 ⊂ Ω2,

Ω1 ̸= Ω2), with a maximum set (maxK) and a minimum set

(minK).

LetK = {Ω𝑡}𝑡∈𝑇be a convex stratification. If for every

𝑠 ∈ 𝑇 \ {max 𝑇} the property ⋂

𝑡>𝑠 Ω𝑡= Ω𝑠 (12)

holds, then as in [1],K = {Ω𝑡}𝑡∈𝑇will be called a quasiconvex

family.

An important quasiconvex family associated with a

con-tinuous self-distancing curve from𝐾, 𝛾: 𝑡 → 𝑥(𝑡) is K =

𝑡}𝑡∈𝑇, where

Ω𝑡= co (𝛾𝑥(𝑡)∪ 𝐾) . (13)

The couple(𝛾, K) is special case of Expanding Couple, a class

introduced in [8].

Remark 6. If𝛾 ∈ SDC𝐾, then for all𝑥 ∈ 𝛾 the curve (𝛾 \ 𝛾𝑥) ∪ {𝑥} is a self-distancing curve from the convex hull of the set

𝛾𝑥∪ 𝐾.

This fact is a direct consequence of the following.

Proposition 7 (see [8, Lemma 4.9]). Let 𝑝, 𝑞, 𝑦𝑖 ∈ R𝑛,𝑖 =

1, . . . , 𝑠. If

󵄨󵄨󵄨󵄨𝑝 − 𝑦󵄨󵄨󵄨󵄨 ≤ 󵄨󵄨󵄨󵄨𝑞 − 𝑦󵄨󵄨󵄨󵄨, 𝑓𝑜𝑟 𝑦 = 𝑦𝑖, 𝑖 = 1, . . . , 𝑠 (14)

then the same holds for every𝑦 ∈ co({𝑦𝑖, 𝑖 = 1, . . . , 𝑠}).

The statement of the previous proposition holds if large inequalities are replaced by strict inequalities everywhere.

2.2. The Support Function of a Plane Convex Body. Let𝐾 ⊂

R𝑛be a convex body not reduced to a point.

For a convex body𝐾, the support function is defined as

𝐻𝐾(𝑥) = sup

𝑦∈𝐾⟨𝑥, 𝑦⟩ , 𝑥 ∈ R

𝑛,

(15)

where⟨⋅, ⋅⟩ denotes the scalar product in R𝑛. For𝑛 = 2, 𝜗 ∈ R,

let𝜃 = (cos 𝜗, sin 𝜗) ∈ 𝑆1 andℎ𝐾(𝜗) fl 𝐻𝐾(𝜃); it will be

denotedℎ(𝜗) if no ambiguity arises.

For every𝜃 ∈ 𝑆1there exists at least one point𝑥 ∈ 𝜕𝐾

such that

⟨𝜃, 𝑦 − 𝑥⟩ ≤ 0 ∀𝑦 ∈ 𝐾; (16)

this means that the line through𝑥 orthogonal to 𝜃 supports

𝐾. For every 𝑥 ∈ 𝜕𝐾 let ̂𝑁𝑥be the set of𝜃 ∈ 𝑆1such that (16)

holds. Let𝐹(𝜃) be the set of all 𝑥 ∈ 𝜕𝐾 satisfying (16). If 𝜕𝐾

is strictly convex at the direction𝜃 then 𝐹(𝜃) reduces to one

point and it will be denoted by𝑥(𝜃).

Definition 8. The set valued map𝐺 : 𝜕𝐾 → 𝑆1, 𝜕𝐾 ∋ 𝑥 →

̂

𝑁𝑥 ⊂ 𝑆1, is the generalized Gauss map;𝑥 ∈ 𝜕𝐾 is a vertex on

𝜕𝐾 iff ̂𝑁𝑥is a sector with interior points. The set valued map

𝐹 : 𝑆1 → 𝜕𝐾, 𝑆1 ∋ 𝜃 → 𝐹(𝜃) ⊂ 𝜕𝐾 is the reverse generalized

Gauss map;𝐹(𝜃) is a closed segment, possibly reduced to a

single point, and it will be called 1-face when it has interior points.

Let𝑃 be the covering map

𝑃 : R 󳨀→ 𝑆1, (17)

R ∋ 𝜗 󳨀→

𝜃 = (cos 𝜗, sin 𝜗) ∈ 𝑆1. (18)

Let𝐿 = |𝜕𝐾|, and let 𝑠 → 𝑥𝑙(𝑠), 0 ≤ 𝑠 < 𝐿 (𝑠 → 𝑥𝑟(𝑠),

0 ≤ 𝑠 < 𝐿) be the parametric representations of 𝜕𝐾 depending on the arc length counterclockwise (clockwise) with an initial

point (not necessarily the same). Let us extend𝑥𝑙(⋅) and 𝑥𝑟(⋅)

by defining

𝑥𝑙(𝑠) fl 𝑥𝑙(𝑠 − 𝑘𝐿) if 𝑘𝐿 ≤ 𝑠 < (𝑘 + 1) 𝐿, (𝑘 ∈ Z) , (19)

similarly for𝑥𝑟.

Let us fix𝑥0 ∈ 𝜕𝐾, 𝜃0 ∈ 𝐺(𝑥0), 𝜃0 = (cos 𝜗0, sin 𝜗0), and

𝜗0∈ R.

For later use, we need to have 𝑥0 = 𝑥𝑙(𝑠0) = 𝑥𝑟(𝑠0);

this can be realized by choosing suitable initial points for the

parameterizations𝑥𝑙and𝑥𝑟. Then 𝑥𝑙(𝑠0+ 𝑠) = 𝑥𝑟(𝑠0+ 𝐿 − 𝑠) , ∀𝑠 ∈ R. (20) The maps 𝑥𝑙: R 󳨀→ 𝜕𝐾, 𝑥𝑟: R 󳨀→ 𝜕𝐾 (21)

are covering maps.

The initial parameters will be

𝑥0= 𝑥𝑙(𝑠0) = 𝑥𝑟(𝑠0) ∈ 𝜕𝐾,

𝑆1∋ 𝜃

0∈ 𝐹−1(𝑥0) ,

R ∋ 𝜗0∈ 𝑃−1(𝜃0)

(22)

(𝐹−1(𝑥0), 𝑃−1(𝜃0) are the back images of 𝐹, 𝑃, resp.). Let 𝑘 ∈

Z. Let us define, for 𝜗0+ 2𝑘𝜋 < 𝜗 < 𝜗0+ 2(𝑘 + 1)𝜋:

𝑠𝑙+(𝜗) fl sup {𝑠 ∈ R : 𝑘𝐿 < 𝑠 ≤ (𝑘 + 1) 𝐿, 𝑥𝑙(𝑠)

(4)

if𝜗 = 𝜗0+ 2𝑘𝜋

𝑠𝑙+(𝜗) fl sup {𝑠 ∈ R : 𝑘𝐿 ≤ 𝑠 < (𝑘 + 1) 𝐿, 𝑥𝑙(𝑠)

∈ 𝐹 (𝑃 (𝜗))} . (24)

Similarly, let us define for𝜗0+ 2(𝑘 − 1)𝜋 < 𝜗 < 𝜗0+ 2𝑘𝜋:

𝑠𝑟−(𝜗) fl inf {𝑠 ∈ R : 𝑘𝐿 ≤ 𝑠 < (𝑘 + 1) 𝐿, 𝑥𝑟(𝑠)

∈ 𝐹 (𝑃 (𝜗))} ; (25)

if𝜗 = 𝜗0+ 2𝑘𝜋

𝑠𝑟−(𝜗) fl inf {𝑠 ∈ R : (𝑘 − 1) 𝐿 < 𝑠 ≤ 𝑘𝐿, 𝑥𝑟(𝑠)

∈ 𝐹 (𝑃 (𝜗))} . (26)

The function𝑠𝑙+is increasing inR and right continuous and

with left limits (so called cadlag function). Similar properties

hold for−𝑠𝑟−. Let us recall that a cadlag increasing function

𝑠(𝜗), 𝜗 ∈ R, has a right continuous inverse defined as

𝜗 (𝑠) = inf {𝜗 : 𝑠 (𝜗) > 𝑠} . (27)

Let𝜗𝑙+(⋅) be the right continuous inverse of 𝑠𝑙+(⋅). Let 𝑠 →

𝜗𝑟−(𝑠) be the opposite of the right continuous inverse of

−𝑠𝑟−(⋅).

Let us introduce for simplicity

n𝜗fl (cos 𝜗, sin 𝜗) ,

t𝜗fl (− sin 𝜗, cos 𝜗) . (28)

Let𝜗 → ℎ(𝜗) be the support function of 𝐾.

It is well known ([13]) that if𝜕𝐾 is 𝐶2+(i.e.,𝜕𝐾 ∈ 𝐶2, with

positive curvature), thenℎ is 𝐶2 and the counterclockwise

element arc𝑑𝑠 of 𝜕𝐾 is given by

𝑑𝑠 = (ℎ + ̈ℎ) 𝑑𝜗. (29)

ℎ(𝜗) + ̈ℎ(𝜗) is the positive radius of curvature; moreover the

reverse Gauss map𝐹 : 𝜃 → 𝑥 ∈ 𝜕𝐾 is a 1-1 map given by

𝑥 (𝜃) fl ℎ (𝜗) n𝜗+ ̇ℎ (𝜗) t𝜗, 𝜗 ∈ 𝑃−1(𝜃) . (30)

The previous formula also holds for an arbitrary convex

body, for every𝜗 such that 𝐹(𝜃) is reduced to a point; see [10].

Let us recall that a real valued function𝑥 → 𝑓(𝑥) is called

semiconvex onR when there exists a positive constant 𝐶 such

that𝑓(𝑥) + 𝐶𝑥2is convex onR. From (29) the function 𝜗 →

ℎ(𝜗) + (1/2)𝜗2maxℎ is convex on R; thus ℎ is semiconvex. In

the case that𝐾 is an arbitrary convex body, by approximation

arguments with𝐶2+convex bodies (see [11]) it follows that the

support function of𝐾 is also semiconvex. As consequence ℎ

is Lipschitz continuous, it has left (right) derivative ̇ℎ(resp.,

̇ℎ+) at each point, which is left (right) continuous. Moreover

at each point the right limit of ̇ℎis ̇ℎ+and the left limit of ̇ℎ+

is ̇ℎ; see [14, pp. 228].

It is not difficult to show (from (30), with a right limit

argument) that, for an arbitrary convex body, for𝜗 ∈ R, the

formula

𝑥𝑙(𝑠𝑙+(𝜗)) = ℎ (𝜗) n𝜗+ ̇ℎ+(𝜗) t𝜗 (31)

holds. Similarly the formula

𝑥𝑟(𝑠𝑟−(𝜗)) = ℎ (𝜗) n𝜗+ ̇ℎ(𝜗) t𝜗 (32)

holds.

If𝜕𝐾 is not strictly convex at the direction 𝜃 = (cos 𝜗,

sin𝜗) then ℎ is not differentiable at 𝜗 and

̇ℎ

+(𝜗) − ̇ℎ−(𝜗) = 󵄨󵄨󵄨󵄨𝑥𝑙(𝑠𝑙+(𝜗)) − 𝑥𝑟(𝑠𝑟−(𝜗))󵄨󵄨󵄨󵄨 = |𝐹 (𝜃)| . (33)

If𝑥1, 𝑥2 ∈ 𝜕𝐾 let us define arc+(𝑥1, 𝑥2) as the set of points

of𝜕𝐾 between 𝑥1and𝑥2according to the counterclockwise

orientation of 𝜕𝐾 and arc−(𝑥1, 𝑥2) as the set of points

between𝑥1 and𝑥2, according to the clockwise orientation;

|arc±(𝑥

1, 𝑥2)| denote their length.

Remark 9. It is well known that a sequence of convex body

𝐾(𝑛)converges uniformly to𝐾 if and only if the

correspond-ing sequence of support functions converges in the uniform norm; see [11, pp. 66]. Moreover as the two sequences of the end points of a closed counterclockwise oriented arc of

𝜕𝐾(𝑛)converge, then the sequence of the corresponding arcs

converges to a connected arc of𝜕𝐾 and the sequence of the

corresponding lengths converges too.

Proposition 10. Let 𝐾 be a convex body and ℎ its support

function; then 𝑠𝑙+(𝜗) − 𝑠𝑙+(𝜗0) = ∫ 𝜗 𝜗0 ℎ (𝜏) 𝑑𝜏 + ( ̇ℎ+(𝜗) − ̇ℎ+(𝜗0)) , ∀𝜗 ≥ 𝜗0; (34) 𝑠𝑟−(𝜗0) − 𝑠𝑟−(𝜗) = ∫𝜗 𝜗0 ℎ (𝜏) 𝑑𝜏 + ( ̇ℎ−(𝜗) − ̇ℎ−(𝜗0)) , ∀𝜗 ≤ 𝜗0. (35)

Proof. For every convex body𝐾 not reduced to a point the

function𝜗 → 𝑠𝑙+(𝜗) is defined everywhere and satisfies the

weak form of (29); namely, − ∫

R𝑠𝑙+(𝜂) ̇𝜙 (𝜂) 𝑑𝜂 = ∫R(𝜙 + ̈𝜙) (𝜂) ℎ (𝜂) 𝑑𝜂,

∀𝜙 ∈ 𝐶∞0 (R) .

(36)

Using the fact that 𝜗 → ℎ(𝜗) is Lipschitz continuous,

integrating by parts (36), the formula − ∫ R𝑠𝑙+(𝜂) ̇𝜙 (𝜂) 𝑑𝜂 = − ∫ R ̇𝜙 (𝜂) (∫ 𝜂 0 ℎ (𝜏) 𝑑𝜏 + ̇ℎ (𝜂)) 𝑑𝜂, ∀𝜙 ∈ 𝐶∞0 (R) (37)

(5)

holds. Thus

𝑠𝑙+(𝜂) = 𝑐 + ∫𝜂

0 ℎ (𝜏) 𝑑𝜏 + ̇ℎ (𝜂) , a.e. (38)

with𝑐 constant. Passing to the right limit, the equality

𝑠𝑙+(𝜂) = 𝑐 + ∫𝜂

0 ℎ (𝜏) 𝑑𝜏 + ̇ℎ+(𝜂) , ∀𝜂 ∈ R (39)

holds. Formula (34) follows, by computing𝑠𝑙+(𝜗) − 𝑠𝑙+(𝜗0),

using the previous equality. Similarly (35) is proved.

3. Involutes of a Closed Convex Curve

Definition 11. Let𝐼 be an interval. A plane curve 𝐼 ∋ 𝑡 → 𝑥(𝑡)

is convex if at every point𝑥 it has right tangent vector 𝑇+(𝑥)

and arg𝑇+(𝑥(𝑡)) is not decreasing function.

Let 𝑠 → 𝑥(𝑠) be the arc length parameterization of a

smooth curve; the classical definition of involute starting at

a point𝑥0= 𝑥(𝑠0) of the curve 𝑥(⋅) is

𝑖 (𝑠) = 𝑥 (𝑠) − (𝑠 − 𝑠0) 𝑥󸀠(𝑠) 𝑠 ≥ 𝑠0. (40)

Let us notice that𝑠 is the arc length of the curve, not of

the involute; if𝑠0 = 0, then the starting point of the involute

coincides with the starting point of the curve. It is easy to construct an involute of a convex polygonal line (even if classical definition (40) does not work) by using arcs of circle centered at its corner points; moreover the involute depends on the orientation of the curve.

In this section, involutes for the boundary of an arbitrary

plane convex body𝐾, not reduced to a point, will be defined.

The assumption that𝐾 is an arbitrary convex body is needed

to work with the involutes of the convex sets, not smooth, studied in Section 4.

Let𝐾 ∈ 𝐶2+; let𝑥0be a fixed point of𝜕𝐾; 𝑠 → 𝑥(𝑠) can be

the clockwise parameterization of𝜕𝐾 or the counterclockwise

parameterization. Since there exist two orientations, then two different involutes have to be considered. As noted previously

one can assume that the parameterizations of𝜕𝐾 have been

chosen so that𝑥0= 𝑥𝑙(𝑠0) = 𝑥𝑟(𝑠0).

Definition 12. Let one denote by 𝑖𝑙,𝑥0 the left involute of

𝜕𝐾 starting at 𝑥0 corresponding to the counterclockwise

parameterization of𝜕𝐾 and by 𝑖𝑟,𝑥0the right involute

corre-sponding to the clockwise parameterization. When one needs

to emphasize the dependence on𝐾 of involutes, they will be

written as𝑖𝐾𝑙,𝑥

0,𝑖

𝐾 𝑟,𝑥0.

Remark 13. Let us notice that if𝜌 is a plane reflection with

respect to a fixed axis then

𝑖𝐾𝑟,𝑥0= 𝜌 (𝑖𝜌(𝐾)𝑙,𝜌(𝑥

0)) . (41)

This relation allows us to prove our results for the left invo-lutes only and to state without proof the analogous results for the right involutes.

Theorem 14. Let one fix the initial parameters 𝑥0,𝑠0,𝜃0, and

𝜗0. The left and the right involutes of a plane convex curve starting at𝑥0 ∈ 𝜕𝐾, boundary of a 𝐶2+ plane convex body

𝐾 with support function ℎ, are parameterized by the value 𝜗

related to the outer normaln𝜗to𝐾, as follows:

𝑖𝑙,𝑥0(𝜗) = ℎ (𝜗) n𝜗− (∫ 𝜗 𝜗0 ℎ (𝜏) 𝑑𝜏 − ̇ℎ (𝜗0)) t𝜗, 𝑓𝑜𝑟 𝜗 ≥ 𝜗0, (42) 𝑖𝑟,𝑥0(𝜗) = ℎ (𝜗) n𝜗− (∫ 𝜗 𝜗0 ℎ (𝜏) 𝑑𝜏 − ̇ℎ (𝜗0)) t𝜗, 𝑓𝑜𝑟 𝜗 ≤ 𝜗0. (43)

Proof. In the present case there is a 1-1 mapping between𝜗

and𝑠; from (29), it follows that

𝑠 − 𝑠0= ∫𝜗

𝜗0

ℎ (𝜏) 𝑑𝜏 + ̇ℎ (𝜗) − ̇ℎ (𝜗0) ; (44)

then, changing the variable𝑠 with 𝜗 in (40), with elementary

computation, (42) is obtained (since𝑥󸀠(𝑠) = t𝜗 and (30)

holds). Formula (43) follows from (32) and (35).

For an arbitrary convex body𝐾 in place of (30), formulas

(31) and (32) have to be used.

Definition 15. Let𝐾 be a plane convex body; let

𝑥0= 𝑥 (𝑠0) ∈ 𝜕𝐾,

𝜗+0 fl 𝜗𝑙+(𝑠0) ,

𝑠+0 fl 𝑠𝑙+(𝜗+0) .

(45)

The left involute of𝜕𝐾 starting at 𝑥0will be defined as

𝑖𝑙,𝑥0(𝜗) = 𝑥𝑙(𝑠𝑙+(𝜗)) − (𝑠𝑙+(𝜗) − 𝑠0) t𝜗 for𝜗 ≥ 𝜗+0; (46)

similarly if𝜗−0 := 𝜗𝑟−(𝑠0), 𝑠−0 := 𝑠𝑟−(𝜗−0), the right involute

starting at𝑥0will be defined as

𝑖𝑟,𝑥0(𝜗) = 𝑥𝑟(𝑠𝑟−(𝜗)) + (𝑠𝑟−(𝜗) − 𝑠0) t𝜗 for𝜗 ≤ 𝜗0−. (47)

From (46) and (34) it follows that

𝑖𝑙,𝑥0(𝜗) = ℎ (𝜗) n𝜗− (∫ 𝜗 𝜗+ 0 ℎ (𝜏) 𝑑𝜏 − ̇ℎ+(𝜗+0)) t𝜗 − 󵄨󵄨󵄨󵄨𝑥0− 𝑥𝑙(𝑠+0)󵄨󵄨󵄨󵄨 t𝜗, 𝜗 ≥ 𝜗+0; (48)

similarly from (47), (35) it follows that

𝑖𝑟,𝑥0(𝜗) = ℎ (𝜗) n𝜗− (∫𝜗 𝜗− 0 ℎ (𝜏) 𝑑𝜏 − ̇ℎ−(𝜗−0)) t𝜗 + 󵄨󵄨󵄨󵄨𝑥0− 𝑥𝑟(𝑠−0)󵄨󵄨󵄨󵄨 t𝜗, 𝜗 ≤ 𝜗0−. (49)

Let us notice that in (48) and (49) the same parameter

𝜗 is used, but with different range; it turns out that 𝑖𝑙 is

counterclockwise oriented; instead𝑖𝑟 is clockwise oriented;

𝑥0= 𝑖𝑙,𝑥0(𝜗+

0) = 𝑖𝑟,𝑥0(𝜗

− 0).

(6)

Remark 16. The following facts can be derived from the above

equations:

(i) sinceℎ is Lipschitz continuous for every convex body

𝐾, then the involute 𝑖𝑙,𝑥0is a rectifiable curve;

(ii)𝑖𝑙,𝑥0(𝜗+0) = 𝑥0and

󵄨󵄨󵄨󵄨

󵄨𝑖𝑙,𝑥0(𝜗) − 𝑥𝑙(𝑠𝑙+(𝜗))󵄨󵄨󵄨󵄨󵄨 = 𝑠𝑙+(𝜗) − 𝑠0; (50)

(iii) if𝑥 is a vertex of 𝜕𝐾 then 𝑖𝑙,𝑥0(𝜗), for (cos 𝜗, sin 𝜗) ∈

𝑁𝐾(𝑥), lies on an arc of circle centered at 𝑥 with radius

𝑠𝑙+(𝜗) − 𝑠0;

(iv) the involute (42) satisfies

𝑖𝑙,𝑥0(𝜗 + 2𝜋) = 𝑖𝑙,𝑥0(𝜗) − 𝐿t𝜗, ∀𝜗 ≥ 𝜗

+

0. (51)

Lemma 17. Parameterization (48) of the involute 𝑖𝑙,𝑥0is 1-1 in

the interval[𝜗0+, 𝜗+0 + 2𝜋); moreover, except for at most a finite or countable setF of values 𝜗𝑖,𝑖 = 1, 2, . . . (corresponding to the 1-face𝐹𝜃𝑖of𝜕𝐾), 𝑖𝑙,𝑥0is differentiable and

𝑑

𝑑𝜗𝑖𝑙,𝑥0(𝜗) = (𝑠𝑙+(𝜗) − 𝑠0) n𝜗 𝑓𝑜𝑟 𝜗 > 𝜗

+

0, 𝜗 ∉ F; (52)

furthermore 𝑖𝑙,𝑥0 has left and right derivative with common directionn𝜗at𝜗 = 𝜗𝑖∈ F.

Proof. By differentiating (48) and using (34), equality (52) is

proved. Similar argument, at𝜗 = 𝜗𝑖∈ F, proves that n𝜗is the

common direction of the left and right derivatives.

Remark 18. Let𝜗 → 𝑖𝑙,𝑥1(𝜗), 𝜗 → 𝑖𝑙,𝑥2(𝜗), 𝑥𝑖 = 𝑥(𝑠𝑖), 𝑖 = 1, 2,

be left involutes of𝐾. Since

𝑖𝑙,𝑥2(𝜗) − 𝑖𝑙,𝑥1(𝜗) = (𝑠2− 𝑠1) t𝜗,

for 𝜗 > max {𝜗+𝑙 (𝑠2) , 𝜗+𝑙 (𝑠1)} ,

(53)

then they will be called parallel curves. Moreover, by (51),

𝑖𝑙,𝑥0(𝜗) and 𝑖𝑙,𝑥0(𝜗 + 2𝜋) will also be called parallel.

Theorem 19. If 𝑑󰜚 is the arc element of the involute 𝑖𝑙,𝑥0then

𝜗 → 󰜚(𝜗) is continuous and invertible in 𝜗 ≥ 𝜗+0 with

continuous inverse[0, +∞) ∋ 󰜚 → 𝜗(󰜚). Moreover

𝑑󰜚 = (𝑠𝑙+(𝜗) − 𝑠0) 𝑑𝜗 𝑓𝑜𝑟 𝜗 ≥ 𝜗+0, 𝜗 ∉ F; (54)

the involute is a convex curve with positive curvature a.e.:

𝑑𝜗 𝑑󰜚 = 1 (𝑠𝑙+(𝜗) − 𝑠0) 𝑓𝑜𝑟 𝜗 > 𝜗 + 0, 𝜗 ∉ F, (55) 󰜚 → 𝑖𝑙,𝑥0(𝜗(󰜚)) is 𝐶1everywhere, and 𝑑 𝑑󰜚𝑖𝑙,𝑥0= n𝜗(󰜚). (56)

Moreover the following properties hold.

(i) For every󰜚 > 0 the right derivative

(𝑑𝜗𝑑󰜚)+= 1

𝑠𝑙+(𝜗 (󰜚)) − 𝑠0 (57)

exists everywhere and it is a decreasing cadlag function.

(ii)(𝑑/𝑑󰜚)𝑖𝑙,𝑥0has everywhere right derivative given by

(𝑑2

𝑑󰜚2𝑖𝑙,𝑥0)

+

= − 1

𝑠𝑙+(𝜗 (󰜚)) − 𝑠0t𝜗(󰜚). (58)

Theorem 20. Let 𝐾(𝑛) be a sequence of plane convex bodies

which converges uniformly to𝐾, 𝑥(𝑛) ∈ 𝜕𝐾(𝑛),𝑥(𝑛) → 𝑥0; then the corresponding sequences of left involutes𝑖𝐾𝑙,𝑥(𝑛)(𝑛)converge

uniformly to𝑖𝑙,𝑥0in compact subsets of[𝜗0+, +∞]; moreover the corresponding sequence of their derivatives (with respect to the arc length) converges uniformly to(𝑑/𝑑󰜚)𝑖𝑙,𝑥0.

Proof. By Remark 9 the sequence of functions𝑠𝑛𝑙+converges

to𝑠𝑙+. From (54) the arclengths of the left involutes𝑖𝐾𝑙,𝑥(𝑛)(𝑛)

󰜚(𝑛)(𝜗) = ∫𝜗

𝜗0

(𝑠(𝑛)𝑙+ (𝜗) − 𝑠(𝑛)0 ) 𝑑𝜗 (59)

converges uniformly in compact subsets of[𝜗0+, +∞) to the

arc length󰜚(𝜗) of 𝑖𝑙,𝑥0; from (56) the same fact holds for their

derivatives.

Let us consider the arc of the involute

𝜂 fl {𝑖𝑙,𝑥0(𝜗) : 𝜗0+≤ 𝜗 ≤ 𝜗+0 + 3𝜋/2} (60)

and the set valued map𝐹 (Definition 8). Let

𝑄 = ⋃ 𝜗+ 0≤𝜗≤𝜗+0+3𝜋/2 {𝜆𝐹 (𝜃) + (1 − 𝜆) 𝑖𝑙,𝑥0(𝜗) , 0 ≤ 𝜆 ≤ 1} , 𝜃 = (cos 𝜗, sin 𝜗) , (61)

the union of segments joining the points of 𝜂 with the

corresponding points on𝜕𝐾.

Definition 21. If the tangent sector𝑇(𝑥0) to 𝐾 has an opening

less than or equal to𝜋/2 as in Figure 1, then 𝑄 ∪ 𝐾 is convex;

let one define

𝜗∗𝑙 = 𝜗+0 + 3𝜋/2. (62)

If𝑄 ∪ 𝐾 is not convex then let us consider co(𝑄 ∪ 𝐾). Let us

notice that𝜕 co(𝑄 ∪ 𝐾) \ 𝜕(𝑄 ∪ 𝐾) is an open segment with

end points𝐴 , 𝐵, with 𝐴 ∈ 𝜂, 𝐵 ∈ 𝜕𝐾. Let us define 𝜗𝑙∗, with

𝜗0+ + 3𝜋/2 ≤ 𝜗𝑙∗ < 𝜗+0 + 2𝜋 such that (see Figure 2) 𝜃∗𝑙 =

(cos 𝜗𝑙∗, sin 𝜗𝑙∗) is orthogonal to 𝐴𝐵, 𝐵 ∈ 𝐹(𝜃𝑙∗). Let 𝜗1,𝑙 be

the smallest𝜃 > 𝜃0+ satisfying𝐴 = 𝑖𝑙,𝑥0(𝜗1,𝑙). Clearly 𝜗∗𝑙 =

(7)

K x0 𝜃+ 0 𝜃− 0 il,x0(𝜗∗l)

Figure 1: Left involute of a square.

For the right involutes a value𝜗𝑟∗is defined similarly, with

𝜗−

0 − 2𝜋 < 𝜗𝑟∗ ≤ 𝜗0−− 3𝜋/2, such that the line orthogonal to

𝜃∗

𝑟 supporting𝐾 at 𝐹(𝜃∗𝑟) is tangent to the right involute at

𝑖𝑟,𝑥0(𝜗1,𝑟) (see Figure 3) where 𝐹(𝜃∗

𝑟) is the point 𝑥(𝑠𝑟−(𝜗∗𝑟)),

written as𝑥(𝜗∗𝑟) for short.

Theorem 22. Let 𝑖𝑙:= 𝑖𝑙,𝑥0be the left involute starting at𝑥0on the boundary of a plane convex body𝐾; then

(i) the left involute𝜗 → 𝑖𝑙(𝜗) has the distancing from 𝐾

property for𝜗 ≥ 𝜗0+but is not SDC for𝜗 ≥ 𝜗∗𝑙;

(ii) the curve𝜗 ∈ [𝜗+0, 𝜗∗𝑙] → 𝑖(𝜗) is SDC;

(iii) for𝑦 ∈ Int(𝐾) the distance function 𝐽𝑦(𝜗) = |𝑖𝑙(𝜗) − 𝑦|

is strictly increasing for𝜗 ≥ 𝜗+0;

(iv) if𝑦 ∈ 𝜕𝐾, then 𝐽𝑦(𝜗) is not decreasing for 𝜗 ≥ 𝜗0+and

(𝑑/𝑑𝜗)𝐽 > 0 for (cos 𝜗, sin 𝜗) ∉ 𝑁𝐾(𝑦).

Proof. As𝑖𝑙is rectifiable, then the function𝐽𝑦2(𝜗) = |𝑖𝑙(𝜗)−𝑦|2

is an absolutely continuous function for𝜗 ≥ 𝜗+0, and from (52)

for𝜗 ∉ F 1 2 𝑑 𝑑𝜗𝐽𝑦2= ⟨𝑑𝜗𝑑 𝑖𝑙, 𝑖𝑙(𝜗) − 𝑦⟩ = ⟨(𝑠𝑙(𝜗) − 𝑠0) n𝜗, 𝑥𝑙(𝑠𝑙+(𝜗)) + (𝑠𝑙+(𝜗) − 𝑠0) t𝜗− 𝑦⟩ = (𝑠𝑙+(𝜗) − 𝑠0) ⟨n𝜗, 𝑥𝑙(𝑠𝑙+(𝜗)) − 𝑦⟩ ≥ 0; (63)

the last inequality holds sincen𝜗is the outer normal to𝜕𝐾 at

𝑥𝑙(𝑠𝑙+(𝜗)). Moreover the previous inequality is strict for all 𝜗

if𝑦 ∈ Int(𝐾), and it is also a strict inequality for 𝑦 ∈ 𝜕𝐾 and

𝑦 ∉ 𝐹(𝜃). This proves (iii) and (iv). Then (i) follows from (iii) and Definition 1 of distancing from K property for a curve. To prove (ii) let us recall that SDC satisfies (2); then one has

to prove that the angle at𝑖𝑙(𝜗) between the vector 𝑖𝑙(𝜗) − 𝑖𝑙(𝜏),

𝜗+

0 < 𝜏 < 𝜗 ≤ 𝜗𝑙∗, andn𝜗, the tangent vector at𝑖𝑙(𝜗), is greater

than or equal to𝜋/2; this is equivalent to show that the half

line𝑟𝜗through𝑖𝑙(𝜗) and 𝑥(𝑠𝑙+(𝜗)) orthogonal to n𝜗supports

at𝑖𝑙(𝜗) the arc of 𝑖𝑙from𝑥0to𝑖𝑙(𝜗). By Definition 21 this is the

case for all𝜗 between 𝜗+0 and𝜗𝑙∗.

il,x0(𝜗∗l)

K B

x0

A = il(𝜗1,l)

Figure 2: Left involute of an hexagon.

il(𝜗∗r+ 2𝜋) ir(𝜗1,r) x(𝜗∗ r) x0 x(𝜗1,r) K P+(𝜗∗r+ 2𝜋) x(𝜓(𝜗∗ r+ 2𝜋)) y = il,x0( ) = i𝜗l̃ r,x0( )̃𝜗r

Figure 3: Involutes of a circumference.

Corollary 23. The left involute 𝜗 → 𝑖𝑙,𝑥0(𝜗) of the boundary

of a plane convex body𝐾 is a self-distancing curve from 𝐾 for

𝜗 ∈ [𝜗+0, 𝜗∗𝑙]; similarly right involute (49) is a self-contracting

curve from𝐾 for 𝜗 ∈ [𝜗𝑟∗, 𝜗0].

Proof. From (i) of Theorem 22 the left involute is a curve such

that the distance of its points from all𝑦 ∈ 𝐾 is not decreasing;

(ii) of the same theorem proves that it is a SDC. Let us recall that a self-contracting curve is a self-distancing curve with opposite orientation.

Theorem 24. Let 𝐾 be a plane convex body not reduced to a

single point and let𝑥0, 𝑠0, 𝜃0, 𝜗0be the initial parameters. Let

[𝜗+

0, 𝜗0++ 2𝜋] ∋ 𝜗 → 𝑖𝑙(𝜗) be an arc of the left involute starting

at𝑥0, and let[𝜗−0 − 2𝜋, 𝜗−0] ∋ 𝜗 → 𝑖𝑟(𝜗) be an arc of the right involute ending at𝑥0; then there exists only one point𝑦 ̸= 𝑥0 which belongs to both arcs and

𝑦 = 𝑖𝑙(̃𝜗𝑙) = 𝑖𝑟(̃𝜗𝑟) , (64)

with

𝜗0−≤ 𝜗𝑟∗+ 2𝜋 < ̃𝜗𝑙< 𝜗+0+3𝜋2 ≤ 𝜗∗𝑙,

𝜗𝑟∗≤ 𝜗0−−3𝜋2 < ̃𝜗𝑟 < 𝜗∗𝑙 − 2𝜋 ≤ 𝜗0+.

(65)

Proof. For simplicity, first let us prove the existence of 𝑦

(8)

R ∋ 𝜗 → 𝑥(𝜗) := 𝑥(𝜃) defined by (30) is a parameterization

of𝜕𝐾.

Let𝜗 ∈ [𝜗0, 𝜗0+ 2𝜋] and let 𝑃+(𝜗) be the first common

point of the half line{𝑥(𝜗) + 𝜆t𝜗, 𝜆 > 0} and of 𝑖𝑙. Moreover,

let[𝜗0, 𝜗∗𝑙] ∋ 𝜗 → 𝜓(𝜗) be the function satisfying

𝑃+(𝜗) = 𝑖

𝑙(𝜓 (𝜗)) . (66)

Let

𝜙 (𝜗) fl 󵄨󵄨󵄨󵄨𝑃+(𝜗) − 𝑖𝑙(𝜗)󵄨󵄨󵄨󵄨 . (67)

First the following sentence will be proved.

Claim 1.𝑃+(𝜗𝑙) belongs to 𝑖𝑟iff the equality

𝜙 (𝜗𝑙) = 𝐿 (68)

holds for some𝜗𝑙∈ [𝜗0, 𝜗0+ 2𝜋], 𝐿 = |𝜕𝐾|.

Proof of Claim 1. If (68) holds, then

󵄨󵄨󵄨󵄨 󵄨𝑃+(𝜗𝑙) − 𝑥 (𝜗𝑙)󵄨󵄨󵄨󵄨󵄨 =󵄨𝑃󵄨󵄨󵄨󵄨 +(𝜗𝑙) − 𝑖𝑙(𝜗𝑙)󵄨󵄨󵄨󵄨󵄨 − 󵄨󵄨󵄨󵄨󵄨𝑖𝑙(𝜗𝑙) − 𝑥 (𝜗𝑙)󵄨󵄨󵄨󵄨󵄨 = 𝐿 − 󵄨󵄨󵄨󵄨󵄨arc+(𝑥0, 𝑥 (𝜗𝑙))󵄨󵄨󵄨󵄨󵄨 = 󵄨󵄨󵄨󵄨󵄨arc+(𝑥 (𝜗𝑙) , 𝑥0)󵄨󵄨󵄨󵄨󵄨 = 󵄨󵄨󵄨󵄨󵄨arc−(𝑥0, 𝑥 (𝜗𝑙− 2𝜋))󵄨󵄨󵄨󵄨󵄨. (69) Thus 𝑃+(𝜗𝑙) = 𝑥 (𝜗𝑙) + 󵄨󵄨󵄨󵄨󵄨arc+(𝑥 (𝜗𝑙) , 𝑥0)󵄨󵄨󵄨󵄨󵄨t𝜗 𝑙 = 𝑥 (𝜗𝑙− 2𝜋) + 󵄨󵄨󵄨󵄨󵄨arc−(𝑥0, 𝑥 (𝜗𝑙− 2𝜋))󵄨󵄨󵄨󵄨󵄨t𝜗𝑙−2𝜋 = 𝑖𝑟(𝜗𝑙− 2𝜋) . (70)

Thus𝑃+(𝜗𝑙) is on both arcs of involutes and the other way

around.

Our aim is to prove that there exists𝜗𝑙 ∈ [𝜗0, 𝜗0+ 3𝜋/2]

such that (68) holds. For this goal we prove next Claims 2 and 3.

Claim 2. The following facts hold in[𝜗0, 𝜗∗𝑙]:

(i)𝜓 is continuously differentiable and 𝜓󸀠> 0;

(ii)𝜙󸀠> 0.

Proof of Claim 2. Let us prove thatn𝜗andn𝜓(𝜗)satisfy

⟨n𝜗, n𝜓(𝜗)⟩ < 0. (71)

Let us consider the triangle with vertices 𝑥(𝜗), 𝑖𝑙(𝜓(𝜗)),

𝑥(𝜓(𝜗)). As

󵄨󵄨󵄨󵄨𝑖𝑙(𝜓 (𝜗)) − 𝑥 (𝜓 (𝜗))󵄨󵄨󵄨󵄨 =󵄨󵄨󵄨󵄨arc+(𝑥0, 𝑥 (𝜓 (𝜗)))󵄨󵄨󵄨󵄨

≥ 󵄨󵄨󵄨󵄨arc+(𝑥 (𝜗) , 𝑥 (𝜓 (𝜗)))󵄨󵄨󵄨󵄨 ≥󵄨󵄨󵄨󵄨𝑥(𝜓(𝜗)) − 𝑥(𝜗)󵄨󵄨󵄨󵄨, (72)

the angle between𝑥(𝜓(𝜗)) − 𝑃+(𝜗) and 𝑥(𝜗) − 𝑃+(𝜗) is acute

and the angle between n𝜗 and n𝜓(𝜗) is obtuse. Thus (71)

follows. By definition, 𝜓(𝜗) solves (66); thus 𝜓(𝜗) is the

implicit solution to

⟨𝑖𝑙(𝜓 (𝜗)) − 𝑥 (𝜗) , n𝜗⟩ = 0. (73)

As

𝑑𝜓𝑑 𝑖𝑙(𝜓) , n𝜗⟩ = (𝑠 (𝜓) − 𝑠0) ⟨n𝜓, n𝜗⟩ (74)

is negative by (71), then by Dini’s Theorem equation (73) has

a solution𝜓(𝜃) satisfying

(𝑠 (𝜓) − 𝑠0) ⟨n𝜓, n𝜗⟩ 𝜓󸀠(𝜗) + ⟨𝑖𝑙(𝜓 (𝜗)) − 𝑥 (𝜗) , t𝜗⟩

= 0. (75)

As𝑖𝑙(𝜓(𝜗)) − 𝑥(𝜗) = 𝜆t𝜗(𝜆 > 0) and (71) holds, then 𝜓󸀠> 0,

and𝜓 is strictly increasing and continuously differentiable.

Let us prove (ii). The formula 𝑑 𝑑𝜗󵄨󵄨󵄨󵄨𝑖𝑙(𝜗) − 𝑖𝑙(𝜓 (𝜗))󵄨󵄨󵄨󵄨 2 = 2 ⟨𝑖𝑙(𝜗) − 𝑖𝑙(𝜓 (𝜗)) ,𝑑𝜗𝑑 𝑖𝑙(𝜗) − 𝑑 𝑑𝜗𝑖𝑙(𝜓 (𝜗))⟩ (76)

holds. Let us notice that𝑖𝑙(𝜗) − 𝑖𝑙(𝜓(𝜗)) is parallel to t𝜗; thus

by (52)

⟨𝑖𝑙(𝜗) − 𝑖𝑙(𝜓 (𝜗)) ,𝑑𝜗𝑑 𝑖𝑙(𝜗)⟩ = 0. (77)

On the other hand

− ⟨𝑖𝑙(𝜗) − 𝑖𝑙(𝜓 (𝜗)) ,𝑑𝜗𝑑 𝑖𝑙(𝜓 (𝜗))⟩

= − ⟨−𝑠 (𝜗) t𝜗− 𝜆t𝜗, (𝑠 (𝜓 (𝜗)) − 𝑠0) n𝜓(𝜗)⟩ 𝜓󸀠

= (𝑠 (𝜗) + 𝜆) (𝑠 (𝜓 (𝜗)) − 𝑠0) ⟨t𝜗, n𝜓(𝜗)⟩ 𝜓󸀠.

(78)

As the angle betweent𝜗andn𝜓(𝜗)is acute, then last term in

the above equalities is positive; thus the derivative in the left hand side of (76) is positive and (ii) of Claim 2 follows.

Claim 3. In the interval[𝜗0, 𝜗∗𝑙] the function 𝜙 has values

smaller than𝐿 and greater than 𝐿.

Proof of Claim 3. The angles𝜗∗𝑟 and𝜗1,𝑟have been introduced

in Definition 21. For simplicity𝑥(𝑠𝑟−(𝜗1,𝑟)) will be denoted

with 𝑥(𝜗1,𝑟). Let us consider the convex set bounded by

arc+(𝑥(𝜓(𝜗𝑟∗+ 2𝜋)), 𝑥(𝜗1,𝑟)) and by the polygonal line with

vertices 𝑥(𝜗1,𝑟), 𝑖𝑟(𝜗1,𝑟), 𝑃+(𝜗∗𝑟 + 2𝜋), 𝑥(𝜓(𝜗∗𝑟 + 2𝜋)); see

(9)

Clearly the inequalities 󵄨󵄨󵄨󵄨𝑖𝑟(𝜗1,𝑟) − 𝑃+(𝜗𝑟∗+ 2𝜋)󵄨󵄨󵄨󵄨 < 󵄨󵄨󵄨󵄨𝑖𝑟(𝜗1𝑟) − 𝑥 (𝜗1,𝑟)󵄨󵄨󵄨󵄨 + 󵄨󵄨󵄨󵄨arc−(𝑥 (𝜗 1,𝑟) , 𝑥 (𝜓 (𝜗∗𝑟 + 2𝜋)))󵄨󵄨󵄨󵄨 + 󵄨󵄨󵄨󵄨𝑥 (𝜓 (𝜗∗ 𝑟 + 2𝜋)) − 𝑃+(𝜗𝑟∗+ 2𝜋)󵄨󵄨󵄨󵄨 = 󵄨󵄨󵄨󵄨arc−(𝑥 0, 𝑥 (𝜗1,𝑟))󵄨󵄨󵄨󵄨 + 󵄨󵄨󵄨󵄨arc−(𝑥 (𝜗 1,𝑟) , 𝑥 (𝜓 (𝜗∗𝑟 + 2𝜋)))󵄨󵄨󵄨󵄨 + 󵄨󵄨󵄨󵄨arc−(𝑥 (𝜓 (𝜗∗ 𝑟 + 2𝜋)) , 𝑥0)󵄨󵄨󵄨󵄨 = 𝐿 (79) hold. As 𝜙 (𝜗𝑟+ 2𝜋) = 󵄨󵄨󵄨󵄨𝑖𝑙(𝜗∗𝑟 + 2𝜋) − 𝑃+(𝜗𝑟+ 2𝜋)󵄨󵄨󵄨󵄨 < 󵄨󵄨󵄨󵄨𝑖𝑟(𝜗1,𝑟) − 𝑃+(𝜗∗𝑟 + 2𝜋)󵄨󵄨󵄨󵄨 , (80) using the previous inequalities, one obtains

𝜙 (𝜗𝑟∗+ 2𝜋) < 𝐿. (81)

Let us show now that

𝜙 (𝜗0+3𝜋

2 ) > 𝐿 (82)

holds.

Let𝜌 be the half line with origin 𝑥0and direction−t𝜗0;

𝜌 − {𝑥0} crosses the arc 𝑖𝑟 in a first point𝑦1 = 𝑖𝑟(𝛼1), with

𝛼1< 𝜗0− 𝜋/2. Then

𝑟 fl 󵄨󵄨󵄨󵄨𝑥0− 𝑦1󵄨󵄨󵄨󵄨 < 󵄨󵄨󵄨󵄨𝑦1− 𝑥 (𝛼1)󵄨󵄨󵄨󵄨 +󵄨󵄨󵄨󵄨arc−(𝑥 (𝛼1) , 𝑥0)󵄨󵄨󵄨󵄨

= 𝐿. (83)

The half line𝜌 meets the arc 𝑖𝑙in a point𝑦2and|𝑦2− 𝑥0| = 𝐿.

Property (iii) of Theorem 22 implies that the arc 𝐷 of

the left involute after𝑦2lies outside of the circle centered in

𝑥0and with radius𝐿. Similar property for the right involute

implies that the arc𝐶 of the right involute joining 𝑥0to𝑦1lies

in the circle with center𝑥0and radius𝑟; thus the straight line

tangent to𝐾 at 𝑥(𝜗0+ 3𝜋/2) meets the arc 𝐶 in 𝑖𝑟(𝜗0− 𝜋/2)

and𝐷 in 𝑃+(𝜗0+ 3𝜋/2). Therefore 𝜙 (𝜗0+2𝜋3 ) =󵄨󵄨󵄨󵄨󵄨󵄨󵄨𝑖𝑙(𝜗0+3𝜋2 ) − 𝑃+(𝜗0+3𝜋2 )󵄨󵄨󵄨󵄨󵄨󵄨󵄨 =󵄨󵄨󵄨󵄨󵄨󵄨󵄨𝑖𝑙(𝜗0+3𝜋2 ) − 𝑥 (𝜗0+3𝜋2 )󵄨󵄨󵄨󵄨󵄨󵄨󵄨 +󵄨󵄨󵄨󵄨󵄨󵄨󵄨𝑥(𝜗0+3𝜋 2 ) − 𝑃+(𝜗0+ 3𝜋 2 )󵄨󵄨󵄨󵄨󵄨󵄨󵄨 >󵄨󵄨󵄨󵄨󵄨󵄨󵄨𝑖𝑙(𝜗0+3𝜋 2 ) − 𝑥 (𝜗0+ 3𝜋 2 )󵄨󵄨󵄨󵄨󵄨󵄨󵄨 +󵄨󵄨󵄨󵄨󵄨󵄨󵄨𝑥(𝜗0+3𝜋 2 ) − 𝑖𝑟(𝜗0− 𝜋 2)󵄨󵄨󵄨󵄨󵄨󵄨󵄨 =󵄨󵄨󵄨󵄨󵄨󵄨󵄨arc+(𝑥0, 𝑥 (𝜗0+3𝜋2 ))󵄨󵄨󵄨󵄨󵄨󵄨󵄨 +󵄨󵄨󵄨󵄨󵄨󵄨󵄨arc−(𝑥0, 𝑥 (𝜗0+3𝜋2 ))󵄨󵄨󵄨󵄨󵄨󵄨󵄨 = 𝐿. (84) Inequality (82) is proved.

The intermediate values theorem implies that there exists

𝜗𝑙∈ [2𝜋 + 𝜗∗𝑟, 𝜗0+ 3𝜋/2] such that (68) holds. Claim 1 implies

that

𝑃+(𝜗𝑙) = 𝑖𝑙(𝜓 (𝜗𝑙)) = 𝑖𝑟(𝜗𝑙− 2𝜋) , (85)

so the right involute and the left involute meet each other in

one point and (64) is proved with ̃𝜗𝑙= 𝜓(𝜗𝑙), ̃𝜗𝑟 = 𝜗𝑙− 2𝜋.

By approximation argument the same result holds for an

arbitrary convex body𝐾.

Let us prove now that the point𝑦 is unique. Let us argue

by contradiction. Let𝑃, 𝑄 be two distinct points on 𝑖𝑙∩𝑖𝑟, with

𝑃 ≺ 𝑄 on 𝑖𝑙and𝑖𝑟; then since𝑖𝑙is a distancing curve from𝑥0,

󵄨󵄨󵄨󵄨𝑃 − 𝑥0󵄨󵄨󵄨󵄨 ≤ 󵄨󵄨󵄨󵄨𝑄 − 𝑥0󵄨󵄨󵄨󵄨, (86)

and since𝑖𝑟is a contracting curve to𝑥0,

󵄨󵄨󵄨󵄨𝑃 − 𝑥0󵄨󵄨󵄨󵄨 ≥ 󵄨󵄨󵄨󵄨𝑄 − 𝑥0󵄨󵄨󵄨󵄨, (87)

therefore all the points on the arc of 𝑖𝑙 and of 𝑖𝑟 between

𝑃 and 𝑄 have the same distance from 𝑥0; thus, between𝑃

and𝑄, 𝑖𝑙 and𝑖𝑟 (arc of involutes of a same convex body𝐾)

coincide with the same arc of circle centered at𝑥0; this implies

that𝐾 reduce to the point 𝑥0, which is not possible for the

assumption.

Definition 25. Let𝑧 ∉ 𝐾. Let 𝑧𝑙(𝑧𝑟) ∈ 𝜕𝐾 on the contact set on

the “left” (right) support line to𝐾 through 𝑧. If the contact set

is a 1-face on these support lines, then𝑧𝑙and𝑧𝑟are identified

as the closest ones to𝑧. The triangle 𝑧𝑧𝑙𝑧𝑟is counterclockwise

oriented.

Theorem 26. For every 𝜉 ∈ 𝜕𝐾 let one consider the left

involutes𝑖𝑙,𝜉and the right involutes𝑖𝑟,𝜉parameterized by their arc length󰜚. The maps

𝜕𝐾 × (0, +∞) ∋ (𝜉, 󰜚) 󳨀→ 𝑖𝑙,𝜉(𝜃 (󰜚)) ∈ R2\ 𝐾, 𝜕𝐾 × (0, +∞) ∋ (𝜉, 󰜚) 󳨀→ 𝑖𝑟,𝜉(𝜃 (󰜚)) ∈ R2\ 𝐾 (88) are 1-1 maps.

Proof. Assume, in the proof, that𝑥0 ∈ 𝜕𝐾, 𝜃0 ∈ 𝐺(𝑥0), 𝜗0, 𝑠0

are fixed. Let𝑧 ∉ 𝐾. The tangent sector to the cap body 𝐾𝑧

with vertex z has two maximal segments𝑧𝑧𝑙and𝑧𝑧𝑟on the

sides that do not meet𝐾 (except at the end points 𝑧𝑙and𝑧𝑟).

Let𝜗𝑙be such that𝑧𝑙= 𝑥𝑙(𝑠𝑙+(𝜗𝑙)), and let 𝑠 be such that

󵄨󵄨󵄨󵄨𝑧 − 𝑧𝑙󵄨󵄨󵄨󵄨 = 𝑠𝑙+(𝜗𝑙) − 𝑠. (89)

Let𝜉𝑙= 𝑥𝑙(𝑠); let 𝜗 = 𝜗+𝑙(𝑠). From (50) and from the definition

of left involute (46) (with𝜉𝑙in place of𝑥0,𝜗 in place of 𝜗0+, and

𝑠 in place 𝑠0)

(10)

holds; thus the map(𝜉, 󰜚) → 𝑖𝑙,𝜉(󰜚) is surjective. Moreover the map is also injective, since the left involutes do not cross each other since they are parallel (see Remark 18). Similar proof holds for the right involutes.

Let𝜉𝑙 = 𝑥𝑙(𝑠) be the starting point of the left involute 𝑖𝑙,𝜉𝑙

through𝑧, defined in the previous theorem; similarly let 𝜉𝑟be

the starting point of the right involute𝑖𝑟,𝜉𝑟through𝑧. Let us

notice that𝑖𝑙,𝜉𝑙and𝑖𝑟,𝜉𝑟meet each other in a countable ordered

set of points.

3.1. J-Fence and G-Fence

Definition 27. Let𝐾 be a convex body in R2,|𝜕𝐾| > 0, 𝑥0

𝜕𝐾, 𝜃0∈ 𝐺(𝑥0), 𝜃0= (cos 𝜗0, sin 𝜗0), 𝑠0∈ R. Let 𝑖𝑙fl 𝑖𝑙,𝑥0, and

let𝑖𝑟fl 𝑖𝑟,𝑥0. Let

𝑦 = 𝑖𝑙(̃𝜗𝑙) = 𝑖𝑟(̃𝜗𝑟) ∈ R2\ 𝐾 (91)

be the first point where the two involutes cross each other (see Theorem 24). Let one define

J𝑙(𝐾, 𝑥0) fl {𝑦 ∈ R2: 𝑦 = 𝑡𝑥0+ (1 − 𝑡) 𝑖𝑙(𝜗) , 0 ≤ 𝑡 ≤ 1, 𝜗+0 ≤ 𝜗 ≤ ̃𝜗𝑙} , J𝑟(𝐾, 𝑥0) fl {𝑦 ∈ R2: 𝑦 = 𝑡𝑥0+ (1 − 𝑡) 𝑖𝑟(𝜗) , 0 ≤ 𝑡 ≤ 1, ̃𝜗𝑟≤ 𝜗 ≤ 𝜗−0} , J (𝐾, 𝑥0) fl (J𝑙(𝐾, 𝑥0) ∪ J𝑟(𝐾, 𝑥0)) \ Int (𝐾) . (92)

J(𝐾, 𝑥0) will be called the J-fence of 𝐾 at 𝑥0.

Let us notice thatJ𝑙(𝐾, 𝑥0) and J𝑟(𝐾, 𝑥0) are two convex

bodies in common with the segment𝑥0𝑦 only.

From Theorem 26 the starting point𝜉𝑙(𝜉𝑟) of a left (right)

involute is uniquely determined from any point𝑧 ∉ 𝐾 of

the involute. The arc of the points on the left (right) involute

between the starting point and𝑧 will be denoted by 𝑖𝑧𝑙,𝜉

𝑙(𝑖

𝑧 𝑟,𝜉𝑟)

or𝑖𝑧𝑙 (𝑖𝑧𝑟) for short. For𝑦 ⪯ 𝑤 let us denote with 𝑖𝑦,𝑤𝑙 (𝑖𝑦,𝑤𝑟 ) the

oriented arc of the left (right) involute between𝑦 and 𝑤.

Let us introduce now other regions which are bounded by left and right involutes.

Let us fix the initial parameters𝑥0, 𝑠0, 𝜃0, 𝜗0.

Definition 28. Given𝑧 ∈ R2 \ 𝐾, let 𝑖𝑙 = 𝑖𝑙,𝜉𝑙 (𝑖𝑟 = 𝑖𝑟,𝜉𝑟) be

the left (right) involute through𝑧 with starting point 𝜉𝑙(𝜉𝑟)

and let𝑧𝑙 (𝑧𝑟) ∈ 𝜕𝐾 be as in Definition 25. Let 𝜗𝜉+

𝑙satisfying

𝑥𝑙(𝑠𝑙+(𝜗+𝜉𝑙)) = 𝜉𝑙. Let𝜗𝑙 > 𝜗

+

𝜉𝑙 be the smallest angle for which

𝑥𝑙(𝑠𝑙+(𝜗𝑙)) = 𝑧𝑙. Let one consider the parameterization (46);

let one define

G𝑙(𝐾, 𝑧) fl {𝑡𝑥𝑙(𝑠𝑙+(𝜗)) + (1 − 𝑡) 𝑖𝑙(𝜗) , 0 < 𝑡

< 1, 𝜗+𝜉𝑙 < 𝜗 < 𝜗𝑙} . (93)

If𝑖𝑧𝑙 does not cross the open segment𝑧𝑧𝑙, the regionG𝑙(𝐾, 𝑧)

is an open set bounded by the convex arc of left involute𝑖𝑧𝑙, the

segment𝑧𝑧𝑙, and the convex arc of𝜕𝐾: arc+(𝜉𝑙, 𝑧𝑙); otherwise

let𝑤 be the nearest point to 𝑧 where 𝑖𝑧𝑙 crosses the open

segment𝑧𝑧𝑙; the regionG𝑙(𝐾, 𝑧) is an open set bounded by

the arc𝑖𝑤,𝑧𝑙 , the segment𝑤𝑧, and 𝜕𝐾. Similarly let us define

G𝑟(𝐾, 𝑧).

G𝑙(𝐾, 𝑧), G𝑟(𝐾, 𝑧) are open and bounded sets. Let us

define

G (𝐾, 𝑧) fl Int (cl (G𝑙(𝐾, 𝑧) ∪ G𝑟(𝐾, 𝑧))) . (94)

G(𝐾, 𝑧) is an open, bounded, connected set. G(𝐾, 𝑧) will be

called theG-fence of 𝐾 at 𝑧.

Remark 29. If𝑧 is the first crossing point of 𝑖𝑙and𝑖𝑟and𝜉𝑙=

𝜉𝑟, thenG(𝐾, 𝑧) = Int(J(𝐾, 𝜉𝑙)).

Let us conclude this section with the following result, which follows from Theorem 20.

Theorem 30. Let 𝐾 be limit of a sequence of convex bodies

𝐾(𝑛),𝑥

0= lim 𝑥(𝑛)0 , and𝑥(𝑛)0 ∈ 𝜕𝐾(𝑛). Then

J (𝐾(𝑛), 𝑥(𝑛)0 ) 󳨀→ J (𝐾, 𝑥0) . (95)

Moreover if𝑧 ∉ 𝐾, 𝑧 = lim 𝑧(𝑛),𝑧(𝑛)∉ 𝐾(𝑛), then

cl(G (𝐾(𝑛), 𝑧(𝑛))) 󳨀→ cl (G (𝐾, 𝑧)) . (96)

4. Bounding Regions for SDC in the Plane

Let us assume that𝑥0is the end point of one of the following

sets:

(a) a steepest descent curve𝛾;

(b)𝛾𝐾: a self-distancing curve from a convex body𝐾; see

Definition 3.

The following questions arise: can one extend𝛾, 𝛾𝐾beyond

𝑥0? Which regions delimit that extension? Which regions are

allowed and which are forbidden?

Lemma 31. Let 𝑧 ∈ R2\ 𝐾. If 𝑢 ∈ G

𝑙(𝐾, 𝑧) then the arc 𝑖𝑙𝑢

of the left involute to𝐾 ending at 𝑢 is contained in G𝑙(𝐾, 𝑧). Similarly if𝑢 ∈ G𝑟(𝐾, 𝑧), then 𝑖𝑢𝑟 ⊂ G𝑟(𝐾, 𝑧).

Proof. Since𝑢 ∈ G𝑙(𝐾, 𝑧), by (93) there exist 𝜗𝑙 ∈ (𝜗+𝜉𝑙, 𝜗𝑙),

𝜏 ∈ (0, 1) such that

𝑢 = 𝜏𝑥𝑙(𝑠𝑙+(𝜗𝑙)) + (1 − 𝜏) 𝑖𝑙(𝜗𝑙) . (97)

Then the arc𝑖𝑙𝑢 is parallel to an arc of the left involute 𝑖𝑙

(through𝑧) for 𝜗 ∈ (𝜗+𝜉

𝑙, 𝜗𝑙). Then any left tangent segment to

𝐾 from a point of 𝑖𝑢

𝑙 is contained in the left tangent segment

(11)

Lemma 32. Let 𝑧 ∈ R2\ 𝐾 and let 𝑢 ∈ G

𝑙(𝐾, 𝑧). There are

two possible cases:

(i) if the right involute ending at 𝑢 does not cross the

tangent segment 𝑧𝑙𝑧 or it crosses 𝑧𝑙𝑧 at a point 𝑞 ∈

G𝑙(𝐾, 𝑧), then in both cases 𝑖𝑢

𝑟 ⊂ G𝑙(𝐾, 𝑧);

(ii) if the right involute ending at 𝑢 crosses the tangent

segment 𝑧𝑙𝑧 at a point 𝑞 ∈ 𝑧𝑙𝑧 ∩ 𝜕G𝑙(𝐾, 𝑧), then

𝑖𝑞,𝑢

𝑟 \ {𝑞} ⊂ G𝑙(𝐾, 𝑧).

Proof. Since the starting point 𝜉𝑟(𝑢) of the right involute

ending at𝑢 is on 𝜕𝐾, the distance from 𝜉𝑟(𝑢) to a point of

the left involute𝑖𝑧𝑙 is not decreasing; see (iv) of Theorem 22;

similarly the distance from 𝜉𝑟(𝑢) to a point of 𝑖𝑢𝑟 is not

decreasing. In case (i) the arc𝑖𝑢𝑟has its end points inG𝑙(𝐾, 𝑧)

and by the above distance property it can not cross two times the left involute; then it can not cross the boundary

ofG𝑙(𝐾, 𝑧); therefore 𝑖𝑢𝑟 ⊂ G𝑙(𝐾, 𝑧); similarly in case (ii) the

arc𝑖𝑞,𝑢𝑟 can not cross the boundary ofG𝑙(𝐾, 𝑧) at most in 𝑞;

therefore all the points of this arc, except to that𝑞, belong to

G𝑙(𝐾, 𝑧).

From the previous lemma the following follows.

Theorem 33. Let 𝑧 ∉ 𝐾. The following inclusions hold:

(a) if𝑢 ∈ G𝑙(𝐾, 𝑧), then cl(G𝑙(𝐾, 𝑢)) \ 𝜕𝐾 ⊂ G𝑙(𝐾, 𝑧) ; (98) (b) if𝑢 ∈ G𝑟(𝐾, 𝑧), then cl(G𝑟(𝐾, 𝑢)) \ 𝜕𝐾 ⊂ G𝑟(𝐾, 𝑧) ; (99) (c) if𝑢 ∈ G(𝐾, 𝑧), then cl(G (𝐾, 𝑢)) \ 𝜕𝐾 ⊂ G (𝐾, 𝑧) . (100)

Proof. By Lemma 31 the left involute that boundsG𝑙(𝐾, 𝑢) is

insideG𝑙(𝐾, 𝑧); then (98) is proved. Inclusion (99) is proved

similarly. Let𝑢 ∈ G(𝐾, 𝑧) = Int(cl(G𝑙(𝐾, 𝑧) ∪ G𝑟(𝐾, 𝑧))) and

let us consider𝑢 ∈ G𝑙(𝐾, 𝑧); then in case (i) of Lemma 32

also the open arc of the right involute𝑖𝑢𝑟 is insideG𝑙(𝐾, 𝑧) ⊂

G(𝐾, 𝑧). Besides 𝑖𝑢

𝑙 ⊂ 𝜕G𝑙(𝐾, 𝑢); then (100) is trivial. In case

(ii) of Lemma 32 the open arc𝑖𝑞,𝑢𝑟 is insideG𝑙(𝐾, 𝑧). On the

other hand𝑞 is inside G𝑟(𝐾, 𝑧) and by (99) the arc 𝑖𝑞𝑟 ⊂ 𝑖𝑢𝑟 is in

G𝑟(𝐾, 𝑧) ⊂ G(𝐾, 𝑧). Similar arguments hold if 𝑢 ∈ G𝑟(𝐾, 𝑧).

Then (100) holds in this case too.

Lemma 34. Let 𝑤 ∉ 𝐾. Let 𝜂 be polygonal deleted SDC𝐾with end point𝑦 ∈ G(𝐾, 𝑤). Then

𝜂 ⊂ G (𝐾, 𝑤) , (101)

𝜂 ⊂ cl (G (𝐾, 𝑦)) . (102)

Proof. To prove (101), let us assume, by contradiction, that𝜂

has a point𝑧 ∉ G(𝐾, 𝑤). With no loss of generality it can be

assumed that𝑧 ∈ 𝜕G(𝐾, 𝑤) and

𝜂 \ 𝜂𝑧⊂ G (𝐾, 𝑤) . (103)

Then, 𝑧 is the end point of a segment 𝑧𝑤𝑖, where 𝑤𝑖

G(𝐾, 𝑤) ∩ 𝜂 and 𝑧 ≺ 𝑤𝑖on𝜂. As 𝑧 ∈ 𝜕G(𝐾, 𝑤), then there

exists an involute through𝑧 which is a piece of the boundary

ofG(𝐾, 𝑤) (to fix the ideas it is assumed that it is the left

involute𝑖𝑙). Let us consider𝑧𝑙∈ 𝜕𝐾 so that the tangent vector

t𝑧to𝑖𝑙at𝑧 satisfies

⟨t𝑧, 𝑧 − 𝑧𝑙⟩ = 0. (104)

As𝑤𝑖is inside the orthogonal angle centered in𝑧 with sides

t𝑧and𝑧𝑙− 𝑧, then

⟨𝑤𝑖− 𝑧, 𝑧 − 𝑧𝑙⟩ < 0. (105)

Then as for𝜀 > 0 sufficiently small, 𝑧𝜀 := 𝑧 + 𝜀(𝑤𝑖− 𝑧) ∈ 𝜂𝑤𝑖

and at𝑧𝜀the curve𝜂 has tangent vector 𝑤𝑖− 𝑧 that satisfies

⟨𝑤𝑖− 𝑧, 𝑧𝜀− 𝑧𝑙⟩ < 0, (106)

contradicting the fact that 𝜂𝑤 has the distancing from 𝐾

property (5). This proves (101).

If𝑤𝑛 → 𝑦, with 𝑦 ∈ G(𝐾, 𝑤𝑛), also the inclusions

𝜂 ⊂ cl (G (𝐾, 𝑤𝑛)) (107)

hold. Then (102) is obtained by approximation Theorem 30.

Theorem 35. Let 𝐾 be a convex body and let 𝛾𝐾be SDC𝐾,

𝑤 ∈ 𝛾, 𝑤 ∉ 𝐾. Then

𝛾𝑤𝐾⊂ cl (G (𝐾, 𝑤)) . (108)

Proof. Let us choose a sequence{𝑤𝑛}, 𝑤𝑛∈ 𝛾𝐾, 𝑤𝑛⪯ 𝑤, 𝑤𝑛

𝑤. Let us fix the arc 𝛾𝐾

𝑤𝑛. By [8, Theorem 6.16],𝛾

𝐾

𝑤𝑛is limit of

polygonal SDC𝐾with end point𝑤𝑛. From Lemma 34, these

polygonal SDC𝐾are enclosed in cl(G(𝐾, 𝑤𝑛)); then

𝛾𝑤𝐾𝑛 ⊂ cl (G (𝐾, 𝑤𝑛)) (109)

holds too. Inclusion (108) is now obtained from the limit of the previous inclusions and by the approximation Theorem 30.

Theorem 36. Let 𝐾 be a convex body not reduced to a point.

If𝛾𝐾is a self-distancing curve from𝐾 with starting point 𝑥0

𝜕𝐾, then

𝛾𝐾⊂ cl (R2\ (J (𝐾, 𝑥0) ∪ 𝐾)) . (110)

Proof. Let𝑧 be the first crossing point of the left and right

involutes of𝐾 starting at 𝑥0. Then

Int(J (𝐾, 𝑥0)) = G (𝐾, 𝑧) . (111)

By contradiction, if𝛾𝐾 has a point𝑤 ∈ G(𝐾, 𝑧), then, by

Theorem 35, the following inclusion holds:

(12)

since, by the distancing from𝐾 property, 𝛾𝐾has in common

with𝐾 only the starting point 𝑥0then the inclusion

𝛾𝐾𝑤\ {𝑥0} ⊂ cl (G (𝐾, 𝑤)) \ 𝜕𝐾 (113)

holds too. Moreover by (100) the set cl(G(𝐾, 𝑤)) \ 𝜕𝐾 has

positive distance from theR2\ G(𝐾, 𝑧); then 𝛾𝑤𝐾\ {𝑥0} has

a positive distance fromR2\ G(𝐾, 𝑧) = R2\ Int(J(𝐾, 𝑥0)).

This is in contradiction with𝑥0∈ 𝜕J(𝐾, 𝑥0).

Corollary 37. Let 𝛾 be SDC and let 𝑧1∈ 𝛾; then

𝛾 \ 𝛾𝑧1⊂ cl (R2\ J (co (𝛾

𝑧1) , 𝑧1)) . (114)

Proof. Since𝛾 \ 𝛾𝑧1is a self-distancing curve from co(𝛾𝑧1) and

𝑧1 ∈ 𝜕 co(𝛾𝑧1) (see [8, (i) of Lemma 4.6]), then Theorem 36

applies to𝛾𝐾= 𝛾 \ 𝛾𝑧1with𝐾 = co(𝛾𝑧1).

Definition 38. Let𝛾 be SCD. If 𝑧1, 𝑧 ∈ 𝛾, with 𝑧1⪯ 𝑧 let

𝛾𝑧1,𝑧fl 𝛾𝑧\ 𝛾𝑧1. (115)

For𝑧 ∉ 𝐾, let 𝐾𝑧be the cap body, introduced in (9). Next

theorem shows the principal result on bounding regions for

arcs of SDC𝛾.

Theorem 39. Let 𝐾 be a convex body and let 𝛾 be SDC𝐾. If

𝑧1, 𝑧 ∈ 𝛾, with 𝑧1⪯ 𝑧 then

𝛾𝑧1,𝑧⊂ cl (G (𝐾, 𝑧) \ J (𝐾𝑧1, 𝑧

1)) . (116)

Proof. First let us notice that𝛾𝑧1,𝑧has the distancing from𝐾

and from the set point{𝑧1} property; thus by Proposition 7

it has the distancing from𝐾𝑧1property. Then inclusion (116)

follows from Theorems 35 and 36.

Let us conclude the section with the following inclusion

result forJ-fences.

Theorem 40. Let 𝐾, 𝐻 be two convex bodies not reduced to a

point,𝐾 ⊂ 𝐻. Let 𝑥0∈ 𝜕𝐾 ∩ 𝜕𝐻. Then

J (𝐾, 𝑥0) ⊂ J (𝐻, 𝑥0) . (117)

Proof. The boundary of J(𝐻, 𝑥0) consists of two arcs

of the left and right involutes of 𝐻 starting at 𝑥0. By

Corollary 23 they are SDC𝐻, and then they are SDC𝐾;

therefore by Theorem 36 they cannot intersect the boundary

ofJ(𝐾, 𝑥0).

4.1. Minimally Connecting Plane Steepest Descent Curves.

Given a point 𝑥1 ∉ 𝐾, the segment joining it with its

projection𝑥0on𝜕𝐾 is SDC𝐾which minimally connects the

two points.

This subsection is devoted to consider when it would be

possible to connect a given point𝑥0 on the boundary of a

plane convex body𝐾, with an arbitrarily given point 𝑥1∉ 𝐾,

by using a steepest descent curve𝛾 ∈ SDC𝐾. Let us denote

withΓ𝑥𝐾0,𝑥1the class of the curves𝛾 ∈ SDC𝐾starting at𝑥0and

ending at𝑥1. N Kx0 Pl Br V P yr y∗ r yl x1 ̃iP𝑟 r,x0 ̃iP𝑙 l,x0 Ql

Figure 4: The regions𝑁, 𝐵𝑟, and𝑉 when 𝐾 is a square.

Definition 41. Let𝛾 be SDC with end point 𝑦 and let 𝜂 be SDC

with starting point𝑦; let us denote by 𝛾 ∗ 𝜂 the curve joining

𝛾 with 𝜂 in the natural order, if it is SDC curve.

Theorem 42. Let 𝑥0∈ 𝜕𝐾, 𝑥1∉ 𝐾. Then Γ𝑥𝐾0,𝑥1 ̸= 0 iff

𝑥1∈ cl (R2\ (J (𝐾, 𝑥0)) ∪ 𝐾) . (118)

If (118) holds, there exist at most two𝜂𝑖 ∈ Γ𝑥𝐾

0,𝑥1,𝑖 = 1, 2 such

that the following properties are true:

∀𝛾 ∈ Γ𝑥𝐾0,𝑥1 󳨐⇒ co(𝜂1) ⊂ co (𝛾) 𝑜𝑟 co (𝜂2) ⊂ co (𝛾) , (𝑜𝑟 𝑏𝑜𝑡ℎ), (119) ∀𝛾 ∈ Γ𝑥𝐾0,𝑥1 󳨐⇒ 󵄨󵄨󵄨󵄨𝛾󵄨󵄨󵄨󵄨 ≥ min𝑖=1,2{󵄨󵄨󵄨󵄨𝜂𝑖󵄨󵄨󵄨󵄨}. (120) Proof. Let𝛾 ∈ Γ𝑥𝐾

0,𝑥1. From (110) of Theorem 36, since𝑥1∈ 𝛾,

then (118) follows.

Let us prove now that (118) is sufficient. Let us notice that

R2\ (J(𝐾, 𝑥0) ∪ 𝐾) can be divided into four regions 𝑁, 𝐵𝑙,

𝐵𝑟, and𝑉 (see Figure 4) defined as follows:

(i) the closed normal sector𝑁 := 𝑥0+𝑁𝐾(𝑥0) is the angle

bounded by the two half lines𝑡𝑙, 𝑡𝑟 tangent at𝑥0

𝜕𝐾 to the left and right involute 𝑖𝑙 := 𝑖𝑙,𝑥0,𝑖𝑟 := 𝑖𝑟,𝑥0,

respectively; this angle can be reduced to an half line,

starting at𝑥0;

(ii) let𝑃 be the first crossing point between 𝑖𝑙and𝑖𝑟; see

Theorem 24;𝑖𝑙is SDC to𝑖𝑙(𝜗∗𝑙), which will be a point

𝑄𝑙following𝑃; after 𝑄𝑙the involute𝑖𝑙is no more SDC;

see (i) of Theorem 22.

Let us change𝑖𝑙after𝑄𝑙with𝑗𝑙,𝑄𝑙, the left involute of

co(𝐾 ∪ 𝑖𝑄𝑙

𝑙 ) at 𝑄𝑙.

Let us define𝑃𝑙as the first intersection point of𝑗𝑙,𝑄𝑙

with𝜕𝑁, and let

̃𝑖𝑃𝑙 𝑙,𝑥0fl 𝑖 𝑄𝑙 𝑙,𝑥0∗ 𝑗 𝑃𝑙 𝑙,𝑄𝑙. (121)

Riferimenti

Documenti correlati

Dopo la battuta di arresto del 1999 la prima occasione di rilievo per vedere riconosciuto il patrimonio linguistico di rom e sinti come parte della storia e della cultura

34 I libri di viaggio e le relazioni di pellegrinaggio in Terra Santa si scrivono nei volgari ita- liani a partire dal Trecento (i rari testi duecenteschi, come quello toscano edito

We recall that a CAT (constant amortized time) algorithm for the exhaustive generation of parallelogram polyominoes of size n has been recently proposed in [15], where it is shown

A plane is the locus of points determined by a single linear equation, and is parametrized by two free variables.. We begin with the second point

Given a specific instance of the spare parts configuration problem, the application returns the Pareto optimal frontier; its first point corresponds to an empty warehouse

In particular, analysis of NGS data is classified as primary, secondary and tertiary (see Figure 3): primary analysis is essentially responsible of producing raw data;

The MP2 ectodomain with a GPI-anchor had similar stimulatory effects on GHSR1a signalling as full-length MP2, whereas the MP1 ectodomain with MP2 transmembrane and

In conclusione, per affrontare il compito della formazione degli ope- ratori della giustizia penale, tenendo conto di quello che sappiamo sulla formazione nelle