• Nenhum resultado encontrado

A Shape-Newton approach to the problem of covering with identical balls∗

N/A
N/A
Protected

Academic year: 2022

Share "A Shape-Newton approach to the problem of covering with identical balls∗"

Copied!
40
0
0

Texto

(1)

A Shape-Newton approach to the problem of covering with identical balls

Ernesto G. Birgin Antoine Laurain Rafael Massambone Arthur G. Santana June 4, 2021

Abstract

The problem of covering a region of the plane with a fixed number of minimum-radius iden- tical balls is studied in the present work. An explicit construction of bi-Lipschitz mappings is provided to model small perturbations of the union of balls. This allows us to obtain ana- lytical expressions for first- and second-order derivatives using nonsmooth shape optimization techniques under appropriate regularity assumptions. Singular cases are also studied using asymptotic analysis. For the case of regions given by the union of disjoint convex polygons, algorithms based on Voronoi diagrams that do not rely on approximations are given to compute the derivatives. Extensive numerical experiments illustrate the capabilities and limitations of the introduced approach.

Keywords: covering problem, nonsmooth shape optimization, Augmented Lagrangian, New- ton’s method.

AMS subject classification: 49Q10, 49J52, 49Q12

1 Introduction

The problem of covering a region of the plane with a fixed number of minimum-radius identical balls is studied in the present work by expanding the nonsmooth shape optimization approach introduced in [6]. The main challenge in this previous work was the the first-order shape sensitivity analysis with respect to perturbations of the balls’ centers and radii. Therefore, investigating the second-order shape sensitivity is a natural albeit challenging extension of [6].

Shape optimization is the study of optimization problems where the variable is a geometric object; see [13, 17, 34]. One of the key concepts in this discipline is the notion ofshape derivative, that measures the sensitivity of functions with respect to perturbations of the geometry. The theoretical study of second-order shape derivatives is a difficult topic in shape optimization. There exists an abundant literature on the shape Hessian in the smooth setting [10, 11, 13, 34]; while in the

This work has been partially supported by FAPESP (grants 2013/07375-0, 2016/01860-1, 2018/24293-0, and 2019/25258-7) and CNPq (grants 302682/2019-8, 304258/2018-0, and 408175/2018-4).

Department of Computer Science, Institute of Mathematics and Statistics, University of S˜ao Paulo, Rua do Mat˜ao, 1010, Cidade Universit´aria, 05508-090, S˜ao Paulo, SP, Brazil. e-mails: egbirgin@ime.usp.br, rmassam- bone@ime.usp.br, and ags@ime.usp.br

Department of Applied Mathematics, Institute of Mathematics and Statistics, University of S˜ao Paulo, Rua do Mat˜ao, 1010, Cidade Universit´aria, 05508-090, S˜ao Paulo, SP, Brazil. e-mail: laurain@ime.usp.br

(2)

nonsmooth setting it is still an active research topic [24, 25]. Numerical methods based on second- order shape derivative are rarely used in shape optimization due to several difficulties. First of all, the second-order shape derivative is often difficult to compute and costly to implement numerically, especially when partial differential equations are involved. Second, the shape Hessian presents several theoretical issues, such as thetwo norms-discrepancy and lack of coercivity, that have been extensively studied in control problems; see [1, 10] and the references therein. There exist only few attempts at defining numerical methods based on second-order information in shape optimization.

In [14], a regularized shape-Newton method is introduced to solve an inverse problem for star- shaped geometries. Second order preconditioning of the shape gradient has been used in [19] for image segmentation and in [3, 32] for aerodynamic optimization. Automatic shape differentiation has also been successfully employed to compute first- and second-order shape derivatives [16, 31].

We also observe that the numerical investigations using Newton-type algorithms [14, 19] are set in a relatively smooth setting. In [25], the shape Hessian was calculated for nonsmooth geometries and polygons in a form that was convenient for numerical experiments, but no numerical investigations were performed. To the best of our knowledge, the present paper is the first attempt at designing and analyzing a shape-Newton algorithm in a genuinely nonsmooth setting.

From a theoretical perspective, the main achievement of [6] was to build bi-Lipschitz transfor- mations to model the geometry perturbations corresponding to covering with identical balls. In the present work, these transformations are key elements for the calculation of the second-order shape derivative, which, unlike the first-order shape derivative, differs from the expression that would be obtained in a smooth setting. Indeed, for the piecewise smooth shapes considered in the covering problem, various terms with a support at singular boundary points, typically circles intersection, appear in the shape Hessian.

Due to the generality of the regions to be covered considered in [6], in the presented numerical experiments, the function that measures the covering and its first-order derivatives were approxi- mated with discretization strategies that may by very time consuming if high precision is required.

In the present work, by restricting the region to be covered to be the union of disjoint convex polygons, algorithms based on Voronoi diagrams to compute the covering function and its first- and second-order derivatives without relying on approximations are given.

The problem of covering a two-dimensional region with identical balls has already been consid- ered in the literature. Covering equilateral triangles and squares was considered in [29] and [30], respectively; while covering the union and difference of polygons was considered in [35]. The cover- ing of rectangles, triangles, squares and arbitrary regions was considered in [18], [27], [28] and [37], respectively. However, the problem addressed in [37] actually consists of covering an arbitrary set of points, which is substantially different from the problem of covering an entire region. All of these papers approach the problem as an optimization problem. In [18, 27, 28] a simulated annealing approach with local search in which the centers of the balls are chosen as points on an adaptive mesh is considered. In [29, 30], a discrete rule is used to define the radius; while a BFGS method is used to solve subproblems in which the radius is fixed. A feasible direction method that requires solving a linear programming problem at each iteration was proposed in [35]. None of the men- tioned works addresses the problem in a unified way as a continuous optimization problem, nor do they present first- or second-order derivatives of the functions that define the problem. In [5], the problem of covering an arbitrary region is modeled as a nonlinear semidefinite programming prob- lem using convex algebraic geometry tools. The introduced model describes the covering problem without resorting to discretizations, but it depends on some polynomials of unknown degrees whose

(3)

coefficients are difficult to compute, limiting the applicability of the method.

The rest of this paper is organized as follows. Section 2 presents a formal definition of the prob- lem, the formula for the first-order derivative introduced in [6], and the formula for the second-order derivative being introduced in the present work. Section 3 presents the derivation of the second- order derivatives for non-degenerate cases; while degenerate cases are considered in Section 4.

Algorithms based on Voronoi diagrams for the exact calculation of the covering function and its first- and second-order derivatives are introduced in Section 5. Extensive numerical experiments are given in Section 6. Final considerations are given in Section 7.

Notation: Given x, y ∈ Rn, x·y = x>y ∈ R; while x⊗y = xy> ∈ Rn×n. The divergence of a sufficiently smooth vector field R2 3 (x, y) 7→ V(x, y) = (V1(x, y), V2(x, y)) ∈ R2 is defined by divV := ∂V∂x1 +∂V∂y2, and its Jacobian matrix is denotedDV. Given an open setS∈Rn,S denotes its closure, ∂S=S\S its boundary, and Vol(S) its volume. LetB(xi, r) denote an open ball with centerxi∈R2 and radiusr. For a sufficiently smooth setS ⊂R2S(z) denotes the unitary-norm outwards normal vector to S atz and τS(z) the unitary-norm tangent vector to∂S atz (pointing counter-clockwise). In the particular case S =B(xi, r) we use the simpler notation νi := νB(xi,r) and τi := τB(xi,r), and we have νi(z) = (cosθz,sinθz)> and τi(z) = (−sinθz,cosθz)>, where θz is the angular coordinate ofz−xi. For intersection pointsz∈∂S∩B(xi, r), we also use the notation ν−i(z) :=νS(z).

2 The shape optimization problem

LetA⊂R2 and Ω(x, r) =∪mi=1B(xi, r) withx:={xi}mi=1. We consider the problem of covering A using a fixed number m of identical balls B(xi, r) with minimum radius r, i.e., we are looking for (x, r)∈R2m+1 such thatA⊂Ω(x, r) with minimumr. The problem can be formulated as

Minimize

(x,r)∈R2m+1 r subject to G(x, r) = 0, (1)

where

G(x, r) := Vol(A)−Vol(A∩Ω(x, r)). (2)

Note that G(x, r) = 0 if and only if A ⊂ Ω(x, r) up to a set of zero measure, i.e., when Ω(x, r) covers A.

The derivatives of G can be computed using techniques of shape calculus [13, 17, 26, 25, 34].

In particular it was shown in [6] that, under suitable assumptions,

∇G(x, r) =− Z

A1

ν1(z)dz,· · · , Z

Am

νm(z)dz, Z

∂Ω(x,r)∩A

dz >

, (3)

where

Ai =∂B(xi, r)∩∂Ω(x, r)∩A (4)

fori= 1, . . . , m.

In the present work, we show that

2G(x, r) =

2xG(x, r) ∇2x,rG(x, r)

2x,rG(x, r)>2rG(x, r)

, (5)

(4)

where ∇2xG(x, r) ∈ R2m×2m, ∇2x,rG(x, r) ∈ R2m, and ∇2rG(x, r) = ∂r2G(x, r) ∈ R are described below. Their description is based on the fact that each setAican be represented by a finite number mi ≥0 of arcs of the circle∂B(xi, r). Note that, since (∪mi=1∂B(xi, r))∩∂Ω(x, r) =∂Ω(x, r), by (4),

m

[

i=1

Ai=∂Ω(x, r)∩A, (6)

i.e., the union of all Ai represents a partition of ∂Ω(x, r)∩A; see Figure 1. Each arc in Ai can be represented by a pair of points (v, w), named starting and ending points, in counter-clockwise direction, i.e., such that the angular coordinatesθv andθwofv−xi andw−xi, respectively, satisfy θv ∈ [0,2π) and θw ∈ (θv, θv + 2π]; see Figure 2. If Ai is not a full circle, we denote by Ai the set of pairs (v, w) that represent the arcs in Ai; otherwise, we define Ai = ∅. In addition, if Ai is a full circle, then we set Circle(Ai) equal to true; otherwise, we set Circle(Ai) equal to false.

We say a configuration (x, r) is non-degenerate if, for every i= 1, . . . , m, every (v, w) ∈ Ai, and everyz∈ {v, w}, there exists one and only oneν−i(z) andν−i(z)·τi(z)6= 0. A characterization of non-degenerate configurations, which satisfy Assumptions 1 and 2, is given in the next section.

(a) (b)

Figure 1: (a) represents a region A to be covered and an arbitrary configuration of balls Ω(x, r).

(b) represents, in red, ∂Ω(x, r)∩A. Each Ai corresponds to the red arcs that intersect ∂B(xi, r).

Note that, in this example, most sets Ai contain two or three maximal arcs; and there is only one set Ai with four maximal arcs.

Assuming (x, r) is non-degenerate, we have that ∇2rG(x, r) in (5) is given by

2rG(x, r) =−Per(∂Ω(x, r)∩A)

r −

m

X

i=1

X

(v,w)∈Ai

s|L(z)| −ν−i(z)·νi(z) ν−i(z)·τi(z)

{w v

, (7)

where, for an arbitrary expression Φ(z),JΦ(z)K

wv := Φ(w)−Φ(v), Per(S) denotes the perimeter of the setS, and, for an extremezof an arc represented by (v, w)∈Ai,L(z) ={`∈ {1, . . . , m} \ {i} |z∈

∂B(x`, r)}.

Matrix ∇2xG(x, r) in (5) is given by the 2×2 diagonal blocks

x2ixiG(x, r) = 1 r

Z

Ai

−νi(z)⊗νi(z) +τi(z)⊗τi(z)dz+ X

(v,w)∈Ai

−i(z)·νi(z)

ν−i(z)·τi(z)νi(z)⊗νi(z) {w

v

(8)

(5)

xi

x`

u v

w

A

ν`(u)

ν−`(u) =νA(u) νi(v)

νi(w) ν−i(w) =ν`(w)

ν−i(v) =νA(v)

Figure 2: The setAi =∂B(xi, r)∩Ω(x, r)∩Ais composed of two arcs (in red). Ifz∈∂B(xi, r)∩

∂B(x`, r) for some ` 6= i, as for z = w, then ν−i(z) = ν`(z), while if z ∈ ∂B(xi, r)∩∂A, as for z∈ {u, v}, thenν−i(z) =νA(z).

and the 2×2 off-diagonal blocks

x2ix`G(x, r) = X

v∈Ii`

νi(v)⊗ν`(v)

ν`(v)·τi(v) − X

w∈Oi`

νi(w)⊗ν`(w)

ν`(w)·τi(w) , (9) where Ii` = {v ∈ ∂B(x`, r) | (v,·) ∈ Ai} and Oi` = {w ∈ ∂B(x`, r) | (·, w) ∈ Ai}. (Note that Ii`=Oi`=∅for all`6=iifAi =∅.) Finally, array∇2x,rG(x, r) in (5) is given by the 2-dimensional arrays

x2irG(x, r) =−1 r

Z

Ai

νi(z)dz+ X

(v,w)∈Ai

u v

ν−i(z)·νi(z)

ν−i(z)·τi(z)νi(z)− X

`∈L(z)

νi(z) τi(z)·ν`(z)

}

~

w

v

. (10)

3 Proof of second-order differentiability of G

In this section, we prove that the second-order derivatives ofG, as defined in (2), are given by (5, 7, 8, 9, 10). In [6] we have built appropriate bi-Lipschitz mappingsTtin order to use integration by substitution for the differentiation of G(x+tδx, r) and G(x, r+tδr). Some of the more technical aspects of these constructions were related to the fact that G(x, r) is an area functional, which required defining Tt on Ω(x, r)∩A and on ∂(Ω(x, r) ∩A). Since ∇G only involves boundary integrals that in addition can be decomposed into integrals on arcs, this facilitates the construction of the mappings Tt required for the calculation of ∇2G(x, r), as Tt only needs to be defined on

∂Ω(x, r)∩A.

We consider two types of transformations for the shape sensitivity analysis. First, in the case of fixed radius and center perturbations one needs a mappingTtbetween the reference set∂Ω(x, r)∩A

(6)

and the perturbed set ∂Ω(x+tδx, r)∩A; see Theorem 2. Second, in the case of fixed centers and radius perturbation one needs a mapping Tt between the reference set ∂Ω(x, r)∩A and the perturbed set ∂Ω(x, r+tδr)∩A; see Theorem 3. The shape sensitivity analysis of ∇G is then achieved through integration by substitution using Tt. The construction of these mappings Tt is similar to the constructions in [6]; however the results are presented in a different way as we need specific properties ofTtto compute the derivatives of∇G. One of the main differences with respect to [6] appears in Theorem 2, where one considers a simultaneous perturbations of all the balls’

center, which allows us to simplify the calculations of the Hessian of G. On the one hand,Tt was used in [6] mainly to prove first-order shape differentiability and its unusual structure did not affect the expression of the first-order shape derivative, in the sense that a similar formula would have been obtained in a smooth setting. On the other hand, the expression of the second-order shape derivative of G at a nonsmooth reference domain Ω differs significantly from the expression that would be obtained for a smooth Ω, as it involves terms with a support at singular boundary points of Ω, and the particular structure of Tt now plays an important role in the calculation of those singular terms. This can be understood by considering that, unlike the first-order derivative, the second-order shape derivative depends on the tangential component of ∂tTt|t=0 on the boundary of the reference domain.

In [6], we have described detailed conditions to avoid degenerate situations and we also discussed various examples of such degeneracies and how they may affect the numerical algorithm. In the present paper we use the same conditions to prove second-order differentiability ofG. To summarize, the main issues when studying the differentiability ofGarise when two balls are tangent or exactly superposed, when the boundaries of more than two balls intersect at the same point, or when Ω(x, r) and Aare not compatible in the sense of Definition 1. The role of Assumptions 1 and 2 is to avoid these singular cases, which allows us to prove second-order differentiability of G. We emphasize that these assumptions only exclude a null-measure set of balls’ configurations in R2m+1, and in Section 4 we show via the study of several singular cases that the second-order differentiability ofG fails when these assumptions are not satisfied.

Assumption 1. The centers {xi}mi=1 satisfy kxi−xjk ∈ {0,/ 2r} for all 1 ≤i, j ≤ m, i 6= j and

∂B(xi, r)∩∂B(xj, r)∩∂B(xk, r) =∅ for all1≤i, j, k ≤m with i, j, k pairwise distinct.

Definition 1. Let ω1, ω2 be open subsets of R2. We callω1 and ω2 compatible ifω1∩ω2 6=∅, ω1 andω2 are Lipschitz domains, and the following conditions hold: (i)ω1∩ω2 is a Lipschitz domain;

(ii) ∂ω1∩∂ω2 is finite; (iii) ∂ω1 and ∂ω2 are locally smooth in a neighborhood of ∂ω1∩∂ω2; (iv) τ1(x)·ν2(x) 6= 0 for all x∈∂ω1∩∂ω2, where τ1(x) is a tangent vector to∂ω1 atx and ν2(x) is a normal vector to ∂ω2 atx.

Assumption 2. Sets Ω(x, r) and A are compatible.

We observe that Ω(x, r) is Lipschitz under Assumption 1, and if, in addition, the intersection of∂Ω(x, r) and∂A is empty, then Assumption 2 holds. Hence, in this particular case we can drop Assumption 2 in Theorems 2 and 3.

We also recall the following basic results, which are key ingredients for the calculation of the shape Hessian ofG.

Theorem 1 (Tangential divergence theorem). Let Γ ⊂ R2 be a Ck open curve, k ≥ 2, with a parameterizationγ, and denote(v, w) the starting and ending points ofΓ, respectively, with respect toγ. Letτ be the unitary-norm tangent vector to Γ,ν the unitary-norm normal vector toΓ, andH

(7)

the mean curvature of Γ, with respect to the parameterizationγ. Let F ∈W1,1(Γ,R2)∩C0(Γ,R2), then we have

Z

Γ

divΓ(F) = Z

Γ

HF ·ν+F(w)·τ(w)−F(v)·τ(v) = Z

Γ

HF·ν+JF(z)·τ(z)K

w v, where divΓ(F) := div(F)−DF ν·ν is the tangential divergence of F onΓ.

Proof. The result follows from [33, § 7.2] and [13, Ch. 9,§ 5.5].

Lemma 1 (Integration by substitution for line integrals). Let Γ⊂R2 be a Ck open curve, k≥2, and ν a unitary-norm normal vector toΓ. LetF ∈C0(Γ,R2) and Tt: Γ→Tt(Γ) be a bi-Lipschitz mapping. Then

Z

Tt(Γ)

F(z)dz = Z

Γ

F(Tt(z))ωt, where

ωt(z) :=kM(z, t)ν(z)k (11)

and M(z, t) := det(DTt(z))DTt(z)−> is the cofactor matrix of DTt(z). Furthermore, we have

tωt|t=0 = divΓV with V :=∂tTt|t=0 on Γ. (12) Proof. See [17, Prop. 5.4.3].

3.1 Construction of a perturbation field for center perturbations

Theorem 2 below employs several ideas from [6, Thm. 3.2 & Thm. 3.6]. However, an important difference is that we consider simultaneous center perturbations for all balls instead of just one, which is more convenient for the calculation of ∇2G. Theorem 2 provides an appropriate mapping Tt for the differentiation of ∂xiG(x+tδx, r) that will be used in Sections 3.4 and 3.5 and for the differentiation of∂rG(x+tδx, r) in Section 3.6.

Theorem 2. Suppose that Assumptions 1 and 2 hold. Then there exists t0 >0 such that for all t∈[0, t0]we have the following decomposition

∂Ω(x+tδx, r)∩A=

k¯

[

k=1

Sk(t), (13)

where ¯kis independent oft, Sk(t) are arcs parameterized by an angle aperture[θk,v(t), θk,w(t)], and t7→θk,v(t),t7→θk,w(t) are continuous functions on[0, t0].

Also, for all t ∈ [0, t0] there exists a bi-Lipschitz mapping Tt : ∂Ω(x, r)∩A → R2 satisfying Tt(∂Ω(x, r)∩A) =∂Ω(x+tδx, r)∩A and Tt(Sk(0)) =Sk(t) for all k= 1, . . . ,¯k. Furthermore, we have

V := ∂tTt|t=0 =δxi+∂tξ(0, θ)rτi onSk(0)⊂∂B(xi, r), (14) where ξ is defined in (20) and

V(z) =δxi− νA(z)·δxi

τi(z)·νA(z)τi(z) if z∈∂B(xi, r)∩∂A, (15) V(z) =δxi−ν`(z)·(δxi−δx`)

τi(z)·ν`(z) τi(z) if z∈∂B(xi, r)∩∂B(x`, r), i6=`. (16)

(8)

Proof. The decomposition (13) relies on Assumptions 1 and 2 and is obtained in a similar way as in [6, Thm. 3.2]. Therefore, in this proof we focus on the construction of the mapping Tt. We observe that each extremity of the arcsSk(t) in the decomposition (13) is either a point belonging to∂B(xi+tδxi, r)∩∂A or a point in∂B(xi+tδxi, r)∩∂B(x`+tδx`, r).

We first provide a general formula for the angle ϑ(t), in local polar coordinates with the pole xi +tδxi, describing an intersection point between the circle ∂B(xi +tδxi, r) and ∂A. Let z ∈

∂B(xi, r)∩∂Aand φbe the oriented distance function toA, defined asφ(x) :=d(x, A)−d(x, Ac), where d(x, A) is the distance from x to the set A. Since Ω(x, r) and A are compatible due to Assumption 2, it follows that ∂Ais locally smooth around the points ∂B(xi, r)∩∂A, hence there exists a neighborhood Uz of z such that the restriction of φ to Uz is smooth, φ(x) = 0 and k∇φ(x)k= 1 for all x∈∂A∩Uz.

Let (r, θz) denote the polar coordinates of z, with the pole xi. Introduce the function ψ(t, ϑ) =φ

xi+tδxi+r cosϑ

sinϑ

.

We compute

ϑψ(0, θz) =r

−sinθz cosθz

· ∇φ

xi+r

cosθz sinθz

=rτi(z)· ∇φ(z).

Since Ω(x, r) andAare compatible,B(xi, r) is not tangent to∂Aand usingk∇φ(z)k= 1 we obtain τi(z)· ∇φ(z)6= 0. Thus, we can apply the implicit function theorem and this yields the existence of a smooth function [0, t0]3t7→ ϑ(t) withψ(t, ϑ(t)) = 0 andϑ(0) =θz. We also compute, using

∇φ(z) =k∇φ(z)kνA(z) since φis the oriented distance function to ∂A, ϑ0(0) =−∂tψ(0, ϑ(0))

ϑψ(0, ϑ(0)) =− ∇φ(z)·δxi

i(z)· ∇φ(z) =− νA(z)·δxi

i(z)·νA(z). (17) We now consider the second case of an intersection point in∂B(xi+tδxi, r)∩∂B(xj+tδx`, r), i6=`. Introduce the functions

ψ(t, ϑ) =kζ(t, ϑ)k2−r2 with ζ(t, ϑ) =xi+tδxi−x`−tδx`+r cosϑ

sinϑ

.

Observe thatϑ7→ζ(t, ϑ) is a parameterization of the circle∂B(xi+tδxi, r) in a coordinate system of center x`, which means that the solutions of the equation ψ(t, ϑ) = 0 describe the intersections between∂B(xi+tδxi, r) and ∂B(x`+tδx`, r). We compute ∂ϑψ(0, ϑ) = 2ζ(0, ϑ)·∂ϑζ(0, ϑ) with

ζ(0, ϑ) =xi−x`+r cosϑ

sinϑ

and ∂ϑζ(0, ϑ) =r

−sinϑ cosϑ

.

Now letz∈∂B(xi, r)∩∂B(x`, r) and letθzbe the corresponding angle in a polar coordinate system with polexi. Since Assumption 1 is satisfied, it is easy to see that

ϑψ(0, θz) = 2ζ(0, θz)·∂ϑζ(0, θz)6= 0.

Hence, the implicit function theorem can be applied to (t, ϑ) 7→ ψ(t, ϑ) in a neighbourhood of (0, θz). This yields the existence, for t0 sufficiently small, of a smooth function t7→ ϑ(t) in [0, t0] such thatψ(t, ϑ(t)) = 0 in [0, t0] and ϑ(0) =θz. We also have the derivative

ϑ0(t) =−∂tψ(t, ϑ(t))

ϑψ(t, ϑ(t)) =−ζ(t, ϑ(t))·∂tζ(t, ϑ(t)) ζ(t, ϑ(t))·∂ϑζ(t, ϑ(t)),

(9)

and in particular, using νi = (cosθz,sinθz)> and τi = (−sinθz,cosθz)>, ϑ0(0) =−(xi−x`+rνi)·(δxi−δx`)

(xi−x`+rνi)·(rτi) =−ν`·(δxi−δx`) rν`·τi

. (18)

We are now ready to build the mapping Tt. Let S(t) ⊂ ∂B(xi +tδxi, r) be one of the arcs parameterized by the angle aperture [θv(t), θw(t)] in the decomposition (13); we have dropped the index k for simplicity. Then, θv(t) and θw(t) are given by ϑ(t) with either θz = θv(0) or θzw(0), andϑ(t) either corresponds to an intersection∂B(xi+tδxi, r)∩∂Aor to an intersection

∂B(xi+tδxi, r)∩∂B(x`+tδx`, r). Thus we defineTt on the arcS(0) as Tt(x) :=xi+tδxi+r

cosξ(t, θ) sinξ(t, θ)

with x=xi+r cosθ

sinθ

∈ S(0), (19) where

ξ(t, θ) :=α(t)(θ−θw(0)) +θw(t) for (t, θ)∈[0, t0]×[θv(0), θw(0)] andα(t) := θθw(t)−θv(t)

w(0)−θv(0). (20) The bi-Lipschitz property ofTton ∂Ω(x, r)∩A is obtained as in the proof of [6, Thm. 3.3].

Finally, differentiating in (19) with respect to t and using ξ(0, θ) = θ we get (14). Then (20) yields ξ(t, θv(0)) = θv(t), ξ(t, θw(0)) = θw(t), ∂tξ(0, θv(0)) = θ0a(0), ∂tξ(0, θw(0)) = θb0(0), consequently using (17) we obtain (15) and using (18) we obtain (16).

3.2 Construction of a perturbation field for radius perturbations

Theorem 3 below relies on several ideas from [6, Thm. 3.3 & Thm. 3.8], and provides an appropriate mappingTt for the differentiation of∂rG(x, r+tδr) that will be used in Section 3.3.

Theorem 3. Suppose that Assumptions 1 and 2 hold. Then there exists t0 >0 such that for all t∈[0, t0]we have the following decomposition

∂Ω(x, r+tδr)∩A=

¯k

[

k=1

Sk(t), (21)

where ¯kis independent oft, Sk(t) are arcs parameterized by an angle aperture[θk,v(t), θk,w(t)], and t7→θk,v(t),t7→θk,w(t) are continuous functions on[0, t0].

Also, for all t ∈ [0, t0] there exists a bi-Lipschitz mapping Tt : ∂Ω(x, r)∩A → R2 satisfying Tt(Ω(x, r)∩A) = ∂Ω(x, r+tδr)∩A and Tt(Sk(0)) = Sk(t) for all k = 1, . . . ,k. In addition, we¯ have

V := ∂tTt|t=0 =δrνi+∂tξ(0, θ)rτi onSk(0)⊂∂B(xi, r), (22) where ξ is defined in (20) and

V(z) =δrνi(z)−δrνA(z)·νi(z)

τi(z)·νA(z)τi(z) ifz∈∂B(xi, r)∩∂A, (23) V(z) =δrνi(z) +δr1−ν`(z)·νi(z)

τi(z)·ν`(z) τi(z) if z∈∂B(xi, r)∩∂B(x`, r), i6=`. (24)

(10)

Proof. The proof has the same structure as the proof of Theorem 2, i.e., we separate the two cases of a point belonging to∂B(xi, r+tδr)∩∂Aand a point in ∂B(xi, r+tδr)∩∂B(x`, r+tδr).

The decomposition (21) relies on Assumptions 1 and 2 and is obtained in a similar way as in [6, Thm. 3.2].

First we consider the case of a point in ∂B(xi, r+tδr)∩∂A. We provide a general formula for the angle ϑ(t), in local polar coordinates with pole xi, describing such an intersection point.

Let z∈ ∂B(xi, r)∩∂Aand (r, θz) denote the polar coordinates of z with centerxi. Let φ be the oriented distance function toA defined as in the proof of Theorem 2. Introduce the function

ψ(t, ϑ) =φ

xi+ (r+tδr) cosϑ

sinϑ

.

We compute

ϑψ(0, θz) =r

−sinθz cosθz

· ∇φ

xi+r

cosθz sinθz

=rτi(z)· ∇φ(z).

Since Ω(x, r) and Aare compatible due to Assumption 2, B(xi, r) is not tangent to∂A and using k∇φ(z)k= 1 we obtain τi(z)· ∇φ(z)6= 0. Thus, we can apply the implicit function theorem and this yields the existence of a smooth function [0, t0]3t7→ϑ(t) with ψ(t, ϑ(t)) = 0 andϑ(0) =θz. We also compute the derivative

ϑ0(0) =−∂tψ(0, ϑ(0))

ϑψ(0, ϑ(0)) =−∂tψ(0, ϑ(0))

ϑψ(0, ϑ(0)) =−δr∇φ(z)·νi(z)

i(z)· ∇φ(z) =−δrνA(z)·νi(z)

i(z)·νA(z) , (25) where we have used ∇φ(z) =k∇φ(z)kνA(z) since φis the oriented distance function to∂A.

Now we provide a general formula for the angle ϑ(t), in local polar coordinates with pole xi, describing an intersection point of two circles∂B(xi, r+tδr) and∂B(x`, r+tδr),i6=`. Introduce

ψ(t, ϑ) =kζ(t, ϑ)k2−(r+tδr)2 with ζ(t, ϑ) =xi−x`+ (r+tδr) cosϑ

sinϑ

.

Observe thatϑ7→ζ(t, ϑ) is a parameterization of the circle∂B(xi, r+tδr) in a coordinate system of center x`, which means that the solutions of ψ(t, ϑ) = 0 describe the intersections between

∂B(xi, r+tδr) and ∂B(x`, r+tδr). We compute ∂ϑψ(0, ϑ) = 2ζ(0, ϑ)·∂ϑζ(0, ϑ) with ζ(0, ϑ) =xi−x`+r

cosϑ sinϑ

and ∂ϑζ(0, ϑ) =r

−sinϑ cosϑ

.

Now letz∈∂B(xi, r)∩∂B(x`, r) and letθzbe the corresponding angle in a polar coordinate system with polexi. Since the conditions of Assumption 1 hold, it is easy to see that

ϑψ(0, θz) = 2ζ(0, θz)·∂ϑζ(0, θz)6= 0.

Hence, the implicit function theorem can be applied to (t, ϑ) 7→ ψ(t, ϑ) in a neighbourhood of (0, θz). This yields the existence, for t0 sufficiently small, of a smooth function t7→ ϑ(t) in [0, t0] such thatψ(t, ϑ(t)) = 0 in [0, t0] and ϑ(0) =θz. We also have the derivative

ϑ0(t) =−∂tψ(t, ϑ(t))

ϑψ(t, ϑ(t)) =−ζ(t, ϑ(t))·∂tζ(t, ϑ(t))−(r+tδr)δr ζ(t, ϑ(t))·∂ϑζ(t, ϑ(t)) ,

(11)

and in particular, using νi = (cosθz,sinθz)> and τi = (−sinθz,cosθz)>, ϑ0(0) =−(xi−x`+rνi)·(δrνi)−rδr

(xi−x`+rνi)·(rτi) = δr r

1−ν`·νi

ν`·τi

. (26)

We are now ready to build the mapping Tt. Let S(t) ⊂ ∂B(xi, r+tδr) be one of the arcs parameterized by the angle aperture [θv(t), θw(t)] in the decomposition (21); we have dropped the indexkfor simplicity. Then,θv(t) andθw(t) are given byϑ(t) with eitherθzv(0) orθzw(0), and ϑ(t) either corresponds to a point in ∂B(xi, r+tδr)∩∂A or to a point in ∂B(xi, r+tδr)∩

∂B(x`, r+tδr).

Next, defineξ(t, θ) andα(t) as in (20). Then, forθ∈[θv(0), θw(0)] we haveξ(t, θ)∈[θv(t), θw(t)]

and ξ(t, θ) is a parameterization ofS(t). A pointx∈ S(0) may be parameterized by x=xi+r

cosθ sinθ

, and define Tt(θ) :=xi+ (r+tδr)

cosξ(t, θ) sinξ(t, θ)

. (27)

Writingξ(t, θ) =θ+β(t, θ) withβ(t, θ) := (α(t)−1)(θ−θb(t)), we observe that cosξ(t, θ)

sinξ(t, θ)

=R(xi, β(t, θ)) cosθ

sinθ

=R(xi, β(t, θ))νi,

where R(xi, β(t, θ)) is a rotation matrix of center xi and angle β(t, θ). Also, thanks to θv(0) <

θw(0) < θv(0) + 2π and θ ∈ [θv(0), θw(0)], there exists a smooth bijection θ : A 3 x 7→ θ(x) ∈ [θv(0), θw(0)]. Thus, using (27) we can define the mapping

Tt(x) :=Tt(θ(x)) =x−rνi(x) + (r+tδr)R(xi, β(t, θ(x)))νi(x) for all x∈ S(0)⊂∂B(xi, r). (28) The bi-Lipschitz property ofTton ∂Ω(x, r)∩A can be obtained as in the proof of [6, Thm. 3.3].

Finally, differentiating in (27) with respect to t and using ξ(0, θ) = θ we get (22). Then (20) yields ξ(t, θa(0)) = θa(t), ξ(t, θb(0)) = θb(t), ∂tξ(0, θa(0)) = θ0a(0), ∂tξ(0, θb(0)) = θ0b(0), conse- quently using (25) we obtain (23) and using (26) we obtain (24).

3.3 Second-order derivative of G with respect to the radius

The first-order derivative of Gwith respect to the radius is given by

rG(x, r) =− Z

∂Ω(x,r)∩A

dz,

see (3) and [6, §3.3] for the detailed calculation. As in [6], the calculation is achieved through integration by substitution using the mapping Tt given by Theorem 3, which requires that As- sumption 1 and Assumption 2 hold. According to Theorem 3, there exists a bi-Lipschitz mapping Ttsatisfying Tt(∂Ω(x, r)∩A) =∂Ω(x, r+tδr)∩A, and this yields, using Lemma 1 on each arc of

∂Ω(x, r)∩A,

rG(x, r+tδr) =− Z

∂Ω(x,r+tδr)∩A

dz =− Z

Tt(∂Ω(x,r)∩A)

dz =− Z

∂Ω(x,r)∩A

ωt(z)dz.

(12)

Thus, using Lemma 1 and the decomposition (6), we compute d

dt∂rG(x, r+tδr) t=0

=− Z

∂Ω(x,r)∩A

divΓV(z)dz =−

m

X

i=1

Z

Ai

divΓV(z)dz.

Applying Theorem 1 for each arc inAi, we obtain d

dt∂rG(x, r+tδr) t=0

=−

m

X

i=1

Z

Ai

HV ·νidz−

m

X

i=1

X

(v,w)∈Ai

JV(z)·τi(z)K

wv. (29)

To get a more explicit formula we need to determine V(v), V(w) andV ·νi on Ai. For this we apply Theorem 3 to two different cases. On the one hand, if v ∈ ∂B(xi, r)∩∂B(x`, r) for some i6=`, then applying (24) we obtain

V(v)·τi(v) =δr1−ν`(v)·νi(v)

ν`(v)·τi(v) . (30)

On the other hand, if v∈∂B(xi, r)∩∂A, then applying (23) we get V(v)·τi(v) =−δrνA(v)·νi(v)

τi(v)·νA(v). (31)

Then, recalling that L(z) = {` ∈ {1, . . . , m} \ {i} | z ∈ ∂B(x`, r)} for z ∈ {v, w}, and that ν−i(z) := ν`(z) if z ∈∂B(xi, r)∩∂B(x`, r), `6= i, and ν−i(z) := νA(z) if z ∈∂B(xi, r)∩∂A, we can merge (30) and (31) into a unique formula:

V(v)·τi(v) =δr|L(v)| −ν−i(v)·νi(v)

ν−i(v)·τi(v) . (32)

In a similar way, we also obtain

V(w)·τi(w) =δr|L(w)| −ν−i(w)·νi(w) ν−i(w)·τi(w) . Gathering these results we get

d

dt∂rG(x, r+tδr) t=0

=−δrPer(∂Ω(x, r)∩A)

r −δr

m

X

i=1

X

(v,w)∈Ai

s|L(z)| −ν−i(z)·νi(z) ν−i(z)·τi(z)

{w v

,

where we have used HV ·νi = δrr on Ai ⊂ ∂B(xi, r) due to (22) and H = 1/r. Thus we have obtained (7).

3.4 Second-order derivative of G with respect to the centers The first-order derivative of Gwith respect to the centerxi is given by

xiG(x, r) =− Z

Ai

νi(z)dz.

(13)

see (3) and [6, § 3.4] for the detailed calculation. As in [6], the calculation is achieved through integration by substitution using the mapping Tt from Theorem 2 with the specific perturbation δx = (0, . . . ,0, δxi,0, . . . ,0), which requires that Assumption 1 and Assumption 2 hold. Using Lemma 1 yields

xiG(x+tδx, r) =− Z

∂B(xi+tδxi,r)∩∂Ω(x+tδx,r)∩A

νt(z)dz=− Z

Tt(Ai)

νt(z)dz

=− Z

Ai

νt(Tt(z))ωt(z)dz,

whereνtis the outward unit normal vector to ∂B(xi+tδxi, r)∩∂Ω(x+tδx, r)∩A andωt is given by (11).

To obtain the derivative of ∂xiG(x+tδx, r) with respect to t at t = 0 we need the so-called material derivative of the normal vector given by

d

dtνt(Tt(z))|t=0 =−(DΓV)>νi on Ai,

with V := ∂tTt|t=0; see [Walker, Lemma 5.5, page 99]. Here, DΓV := DV −(DV)νi⊗νi denotes the tangential Jacobian ofV onAi. Then, using (12) we obtain

d

dt∂xiG(x+tδx, r) t=0

=− Z

Ai

−(DΓV)>νiidivΓ(V)dz.

This expression can be further transformed using the following tensor relations:

divΓi⊗V) = divΓ(V)νi+ (DΓνi)V and∇Γ(V ·νi) =DΓνi>V +DΓV>νi on Ai. (33) We show thatDΓνi>V =DΓνiV on Ai. Indeed, letW ∈R2 and denote Vτ and Wτ the tangential components of V and W on Ai. Differentiating νi·νi = 1 onAi we get (DΓνi)>νi = 0 and then

(DΓνi)>V ·W = (DΓνi)>Vτ ·W = (DΓνi)>Vτ·Wτ = (DΓνi)Vτ·Wτ,

where we have used the well-known fact that the second fundamental form (Vτ, Wτ)7→(DΓνi)Vτ·Wτ is symmetric. Further,

(DΓνi)>V ·W = (DΓνi)>Wτ·Vτ = (DΓνi)>W ·V = (DΓνi)V ·W on Ai. (34) Now, using (33), (34) we obtain

d

dt∂xiG(x+tδx, r) t=0

=− Z

Ai

divΓi⊗V)− ∇Γ(V ·νi)dz.

Applying Theorem 1 to the integral of divΓi⊗V) on each arc in Ai we get d

dt∂xiG(x+tδx, r) t=0

=− Z

Ai

H(νi⊗V)·νi− ∇Γ(V ·νi)dz− X

(v,w)∈Ai

J(νi(z)⊗V(z))·τi(z)K

w v,

(14)

and then, usingH= 1/r on Ai, d

dt∂xiG(x+tδx, r) t=0

=−1 r

Z

Ai

(V ·νii−r∇Γ(V ·νi)dz− X

(v,w)∈Ai

J(V(z)·τi(z))νi(z)K

w v. (35) Applying (14) yields V ·νi =δxi·νi on Ai.Considering thatδx`= 0 for `6=isince we use the specific perturbation δx= (0, . . . ,0, δxi,0, . . . ,0), (15) and (16) actually provide the same formula in this particular case:

V(z)·τi(z) =δxi·

τi− ν−i

τi·ν−i

(z)

=−

ν−i·νi ν−i·τiδxi·νi

(z) for z∈ {v, w} and (v, w)∈Ai.

(36)

We also have ∇Γ(V ·νi) =∇Γ(δxi·ν) = (DΓν)>δxi and DΓνi=DΓ

cosθ sinθ

=

Γ(cosθ)>

Γ(sinθ)>

= 1 r

θ(cosθ)τi>

θ(sinθ)τi>

= 1

i⊗τi on Ai. Gathering these results and usingV ·νi =δxi·νi onAi we get

d

dt∂xiG(x+tδx, r) t=0

=−1 r

Z

Ai

(δxi·νii−(τi⊗τi)δxidz+ X

(v,w)∈Ai

−i·νi ν−i·τiνi⊗νi

{w v

δxi,

which yields (8).

3.5 Second order derivative with respect to xi and x` of G

As in Section 3.4 we use the mapping Tt from Theorem 2, which requires that Assumptions 1 and 2 hold, but now with the specific perturbation δx = (0, . . . ,0, δx`,0, . . . ,0). This yields a transformation Tt satisfying in particular Tt(Ai) = ∂B(xi, r)∩∂Ω(x+tδx, r)∩A. Then, using Lemma 1 we obtain

xiG(x+tδx, r) =− Z

∂B(xi,r)∩∂Ω(x+tδx,r)∩A

νt(z)dz=− Z

Tt(Ai)

νt(z)dz=− Z

Ai

νt(Tt(z))ωt(z)dz, where νt is the outward unit normal vector to ∂B(xi, r)∩∂Ω(x+tδx, r)∩A and ωt is given by (11). Applying (14) and considering that δxi = 0 since we are using the specific perturbation δx= (0, . . . ,0, δx`,0, . . . ,0), we get

V ·νi= 0 on Ai. (37)

Then, applying (16) withδxi = 0 we get V(z)·τi(z) = δx`·ν`(z)

τi(z)·ν`(z) ifz∈∂B(x`, r)∩∂B(xi, r), i6=`. (38) Next, the derivative of∂xiG(x+tδx, r) with respect totatt= 0 is already calculated in (35), but the terms (V ·νii and ∇Γ(V ·νi) in (35) vanish due to (37). We also observe thatV(z) = 0

(15)

ifz ∈ {v, w} with (v, w) ∈Ai and z /∈∂B(x`, r). Finally, usingIi` ={v∈∂B(x`, r)|(v,·)∈Ai}, Oi`={w∈∂B(x`, r)|(·, w)∈Ai}and (38) we get

d

dt∂xiG(x+tδx, r) t=0

= X

v∈Ii`

V(v)·τi(v)νi(v)− X

w∈Oi`

V(w)·τi(w)νi(w)

=

 X

v∈Ii`

νi(v)⊗ν`(v)

ν`(v)·τi(v) − X

w∈Oi`

νi(w)⊗ν`(w) ν`(w)·τi(w)

δx`,

which yields (9).

3.6 Second order derivative with respect to xi and r of G

In a similar way as in Sections 3.4 and 3.5, we use the mappingTtfrom Theorem 2 with the specific perturbationδx= (0, . . . ,0, δxi,0, . . . ,0). This yields, using Lemma 1,

rG(x+tδx, r) =− Z

∂Ω(x+tδx,r)∩A

dz=− Z

Tt(∂Ω(x,r)∩A)

dz=− Z

∂Ω(x,r)∩A

ωt(z)dz.

Proceeding as in the calculation leading to (29), we get d

dt∂rG(x+tδx, r) t=0

=−

m

X

`=1

Z

A`

HV ·ν`dz−

m

X

`=1

X

(v,w)∈A`

JV(z)·τ`(z)K

w

v. (39)

Considering thatδx`= 0 for`6=i, since we use the specific perturbationδx= (0, . . . ,0, δxi,0, . . . ,0), (15) and (16) actually provide the same formula in this particular case:

V(z)·τi(z) =δxi·

τi− ν−i

τi·ν−i

(z)

=−

ν−i·νi

ν−i·τi

δxi·νi

(z) for z∈ {v, w} and (v, w)∈Ai,

(40)

and also

V(z)·τ`(z) =δxi·

τ`− ν`

τi·ν`i·τ`)

(z)

=δxi· µ

τi·ν`

(z) if z∈∂B(xi, r)∩∂B(x`, r) and `6=i.

withµ:=τ`i·ν`)−(τi·τ``. This yieldsµ·τi = 0 and

µ·νi= (τ`·νi)(τi·ν`)−(τi·τ`)(ν`·νi) =−(τi·ν`)2−(τi·τ`)2 =−1,

where we have used the geometric properties τ`·νi =−τi·ν` and τi ·τ``·νi. Thus µ =−νi and we get

V(z)·τ`(z) =− δxi·νi(z)

τi(z)·ν`(z) ifz∈∂B(xi, r)∩∂B(x`, r) and`6=i. (41)

(16)

In (39), we observe thatV(z) = 0 wheneverz∈ {v, w}and z /∈∂B(xi, r); this can be seen from (15)-(16) and the fact that we use the specific perturbation δx = (0, . . . ,0, δxi,0, . . . ,0). Hence, recalling that L(z) ={`∈ {1, . . . , m} \ {i} |z∈∂B(x`, r)},

m

X

`=1

X

(v,w)∈A`

JV(z)·τ`(z)K

w

v = X

(v,w)∈Ai

JV(z)·τi(z)K

w v +

m

X

`=1`6=i

X

(v,w)∈A`

JV(z)·τ`(z)K

w v

= X

(v,w)∈Ai

JV(z)·τi(z)K

w

v − X

(v,w)∈Ai

u v

X

`∈L(z)

V(z)·τ`(z) }

~

w

v

.

Note that the negative sign in front of the last sum is due to the fact that if an ending point of an arc in A` belongs to some arc in Ai, then it is a starting point for this arc in Ai, and vice versa.

Using (14) we have V ·ν`≡0 on A` for all `6=i. Since H= 1/r, we may write (39) as d

dt∂rG(x+tδx, r) t=0

=−1 r

Z

Ai

V ·νidz− X

(v,w)∈Ai

u

vV(z)·τi(z)− X

`∈L(z)

V(z)·τ`(z) }

~

w

v

. (42)

Using (14) we getV ·νi =δxi·νi on Ai. Finally, using (40)-(41) we get d

dt∂rG(x+tδx, r) t=0

=−1 r

Z

Ai

δxi·νidz

+ X

(v,w)∈Ai

u v

ν−i(z)·νi(z)

ν−i(z)·τi(z)δxi·νi(z)− X

`∈L(z)

δxi·νi(z) τi(z)·ν`(z)

}

~

w

v

,

which yields (10).

4 Analysis of singular cases

The gradient∇Gand Hessian∇2Gwere obtained under Assumptions 1 and 2, and in this section we investigate several singular cases where these assumptions are not satisfied. On the one hand, it is shown in [6,§ 3.5] that Gis often differentiable even when Assumptions 1 and 2 do not hold, and in the few cases where Gis not differentiable it is at least Gateaux semidifferentiable. On the other hand, Gis never twice differentiable in any of the singular geometric configurations studied in this section. Nevertheless, Gateaux semidifferentiability of the components of∇Gcan often be proven.

We recall here that f :Rn→R is Gateaux semidifferentiable at xin the direction v if

t&0lim

f(x+tv)−f(x)

t exists inRn, and thatf has a derivative in the direction v atx if

t→0lim

f(x+tv)−f(x)

t exists inRn.

Referências

Documentos relacionados

No Brasil, a Lei nº 11.121, de 25/05/2005, também instituiu o 28 de abril como o Dia Nacional em Memória das Vítimas de Acidentes e Doenças do Trabalho , com o objetivo de

This log must identify the roles of any sub-investigator and the person(s) who will be delegated other study- related tasks; such as CRF/EDC entry. Any changes to

Além disso, o Facebook também disponibiliza várias ferramentas exclusivas como a criação de eventos, de publici- dade, fornece aos seus utilizadores milhares de jogos que podem

Jornalismo e Sociedade A Comunicação Estratégica a nível meso (grupos e organizações) e macro (sociedades, mercados) Paula Espírito Santo Comunicação política Media

É uma ramificação da ginástica que possui infinitas possibilidades de movimentos corporais combinados aos elementos de balé e dança teatral, realizados fluentemente em

O credor que não requerer perante o Juízo da execução a(s) adjudicação(ões) do(s) bem(ns) a ser(em) leiloado(s) antes da publicação deste Edital, só poderá adquiri- lo(s)

O fato de que, em 1513, o Papa Leão X intimou Pedro Mártir de Anghiera a lhe fornecer um relatório sobre o assunto, mostra que, agora já na Era dos Descobrimentos, esse mito ganhava

O paciente evoluiu no pós-operatório imediato com progressão de doença em topografia do tumor primário (figura1), sendo encaminhado ao Oncobeda em janeiro de 2017