• Aucun résultat trouvé

Counterexample to regularity in average-distance problem

N/A
N/A
Protected

Academic year: 2022

Partager "Counterexample to regularity in average-distance problem"

Copied!
16
0
0

Texte intégral

(1)

Ann. I. H. Poincaré – AN 31 (2014) 169–184

www.elsevier.com/locate/anihpc

Counterexample to regularity in average-distance problem

Dejan Slepˇcev

Department of Mathematical Sciences, Carnegie Mellon University, Pittsburgh, PA 15213, United States Received 27 September 2012; received in revised form 3 February 2013; accepted 5 February 2013

Available online 19 February 2013

Abstract

The average-distance problem is to find the best way to approximate (or represent) a given measure μ on Rd by a one- dimensional object. In the penalized form the problem can be stated as follows: given a finite, compactly supported, positive Borel measureμ, minimize

E(Σ )= Rd

d(x, Σ ) dμ(x)+λH1(Σ )

among connected closed sets,Σ, whereλ >0,d(x, Σ )is the distance fromxto the setΣ, andH1is the one-dimensional Hausdorff measure. Here we provide, for anyd2, an example of a measureμwith smooth density, and convex, compact support, such that the global minimizer of the functional is a rectifiable curve which is notC1. We also provide a similar example for the constrained form of the average-distance problem.

©2013 Elsevier Masson SAS. All rights reserved.

MSC:49Q20; 49K10; 49Q10; 05C05; 35B65

Keywords:Average-distance problem; Nonlocal variational problem; Regularity

1. Introduction

Given a positive, compactly supported, Borel measureμonRd,d2,λ >0, andΣ a nonempty subset ofRd consider

E(Σ )=

Rd

d(x, Σ ) dμ(x)+λH1(Σ ). (1)

The average-distance problem is to minimize the functional overA= {Σ⊂Rd: Σ− connected and compact}.

E-mail address:slepcev@math.cmu.edu.

0294-1449/$ – see front matter ©2013 Elsevier Masson SAS. All rights reserved.

http://dx.doi.org/10.1016/j.anihpc.2013.02.004

(2)

The problem was introduced by Buttazzo, Oudet, and Stepanov[1]and Buttazzo and Stepanov[2]. They studied the problem in the constrained form, where instead ofH1penalization one minimizes

F (Σ )=

Rd

d(x, Σ ) dμ(x) overA1:=

ΣA: H1(Σ )

. (2)

Over the past few years there has been a significant progress on understanding of the functional; some of which we outline below. An excellent overview article has recently been written by Lemenant[3].

The problem has wide ranging applications. When interpreted as a simplified description of designing the optimal public transportation network then μrepresents the distribution of passengers, andΣ is the network. The desire is to design the network that minimizes the total distance of passengers to the network. Another related problem which can be reduced to the average-distance problem, studied in[1], is when we think of passengers as workers that need to get to their workplace. Then two measures are initially given, the distribution of where workers reside and where they work. Again the goal is to find the optimal network that minimizes the total transportation cost (traveling along the network is for free).

A related interpretation is that of finding the optimal irrigation network (theirrigation problem).

Another interpretation, whose application in a related setting is presently investigated by Laurent and the author, is to find a good one-dimensional representation to a data cloud. Hereμrepresents the distribution of data points.

One wishes to approximate the cloud by a one-dimensional object. The first term in (1) then charges the errors in the approximation, while the second one penalizes the complexity of the representation.

The existence of minimizers ofEfollows from the theorems of Blaschke and Goł ˛ab[2]. In this paper we investigate their regularity. It was shown in[2]that, at least ford=2, the minimizer is topologically a tree made of finitely many simple rectifiable curves which meet at triple junctions (no more that three branches can meet at one point). The authors also show that the minimizer is Ahlfors regular (which was extended to higher dimensions by Paolini and Stepanov[4]), but further regularity of branches remained open. Recently Tilli[5]showed that every compact simple C1,1curve is a minimizer of the average-distance problem (in the constrained form (27)) whereμis the characteristic function of a small tubular neighborhood of the curve. This suggests thatC1,1 is the best regularity for minimizers one can expect (even ifμwere smooth). Further criteria for regularity were established by Lemenant[6].

Due to the presence of theH1term one might expect that, ifμis a measure with smooth density,Σis at leastC1. A recent paper by Buttazzo, Mainini, and Stepanov[7]suggests that this may not be the case, and exhibits a measureμ which is a characteristic function of a set in R2, for which there exists a stationary point ofEwhich has a corner.

Furthermore the results on the blow-up of the problem by Santambrogio and Tilli[8]support the possibility of corners.

Here we prove that minimizers which are not C1are indeed possible. That is for anyd2 provide an example of a measureμwith smooth density for which we prove that the minimizer is a curve which has a corner, and is thus notC1. One of the difficulties in dealing with global energy minimizers is that the functional is not convex. To be able to treat them we introduce constructions and an approximation technique that may be of independent interest.

Our approach is based on approximating the measureμof our interest by particle measuresμn(i.e. the ones that have only atoms). For particle measuresμn the average-distance problem (1) has a discrete formulation that can be carefully analyzed. In particular the minimizers are trees with piecewise linear branches. Our starting point is the construction of a particle measure with three particles,μ, for which we can show that the minimizer is a wedge (curve¯ with exactly two line segments), see Fig.1. We then show that ifμ¯ is smoothed out a bit then the minimizer will still have a corner (even if we also add a smooth background measure of small total mass,q, that makes the support of the perturbed measure convex). We denote the smooth perturbed measure byμq,δwereδis the smoothing parameter.

To show that a minimizer ofEforμq,δ has a corner whenδ andq are small, we consider discrete approximations μq,δ,n of μq,δ. We show that the minimizers Σq,δ,n of E corresponding to μq,δ,n have a corner whose opening is bounded from above independent of n. We furthermore obtain appropriate estimates on the minimizers which guarantee convergence asn→ ∞to a minimizerΣq,δofEcorresponding toμq,δand insure that the corner remains in the limit.

1.1. Outline

In Section 2 we list some of the basic properties of the functional E given in (1), in particular its continuity properties with respect to parameters and scaling with respect to dilation ofμ. In Section3we consider the energy (1)

(3)

withμbeing a particle measure. We obtain conditions for criticality, information of the projection of the measureμ onto the minimizerΣ, and a priori estimates on the curvature (turning angle). The basic three-particle configuration,μ,¯ for which the minimizer is a wedge is also introduced. The construction of the counterexample is carried out in Section 4. We introduce the perturbation and use elementary geometry to obtain various geometric facts about the minimizer of the average-distance problem corresponding to the discrete approximation of the perturbed measure:μq,δ,n. The result of these efforts is that the minimizer must have a corner with a large turning angle (jump in the tangent direction).

Finally we take the limitn→ ∞to obtain that the minimizer forμq,δ has a corner too. In Section5we use a scaling argument to show that the minimizers of the constrained problem (2) can have corners too.

2. Properties of the functional

LetAbe the set of compact connected subsets ofRd. GivenΣA, foryΣ we define theregion of influence ofy,

R(y)=

x∈Rd:(zΣ ) d(x, z)d(x, y)

. (3)

In the next two lemmas we study continuity properties ofEwith respect to dependence onΣandμ.

Lemma 1.For any μPR and anyλ >0, the functionalEμ:A→Ris lower semicontinuous with respect to Hausdorff convergence.

Proof. Assume that ΣnAconverge to Σ in Hausdorff metric, dH. Goł ˛ab’s theorem (see [9]) gives the lower- semicontinuity of theH1measure. Thus it is enough to prove the continuity of the first term of the energy. Note that for anyx∈Rd,

d(x, Σn)d(x, Σ )dHn, Σ ).

To illustrate why, assume that for somex, d(x, Σn) > d(x, Σ )+dHn, Σ ).

Consideringyto be the closest point toxonΣgives d(x, Σn) > d(x, y)+ inf

zΣnd(y, z)d(x, Σn) which is a contradiction. Thus

Rd

d(x, Σn) dμ(x)

Rd

d(x, Σ ) dμ(x)

dHn, Σ )μ Rd

which implies the claim. 2

Lemma 2.ConsiderΣAandλ >0. The mappingμEμ(Σ )is continuous with respect to weak-∗convergence of measures inPR.

Proof. We recall that the Wasserstein metric,dW, metrizes the weak-∗convergence of measures on the set of measures supported inB(0, R). Therefore ifμn μ inPRthendWn, μ)→0 asn→ ∞. Hence there exists a coupling (i.e.

a transportation plan),Πnbetweenμandμnsuch that

B(0,R)×B(0,R)

|xy|2n(x, y)→0 asn→ ∞.

(4)

Therefore

Rd

d(x, Σ ) dμn(x)

Rd

d(x, Σ ) dμ(x) =

Rd×Rd

d(x, Σ )d(y, Σ ) dΠn(x, y)

Rd×Rd

|xy|n(x, y)

R dWn, μ)→0 asn→ ∞. 2 (4)

Lemma 3.Ifμn μ in the weak topology of measures inPRthenEμnΓ Eμwith respect to Hausdorff convergence of sets onA.

Proof. To prove theΓ-convergence we need to show the following

• Lower-semicontinuity. Assumeμn

μandΣnΣin Hausdorff metric asn→ ∞. Then lim inf

n→∞ Eμnn)Eμ(Σ ).

• Construction. Assumeμn μ. For any ΣAthere exists a sequenceΣnA, such thatΣnΣin Hausdorff metric and

nlim→∞Eμnn)=Eμ(Σ ).

The construction claim follows from Lemma2by takingΣn=Σ.

Let us consider the lower-semicontinuity. As before,μn μ impliesdWn, μ)→0 asn→ ∞. As in the estimate (4)

Rd

d(x, Σn) dμn(x)

Rd

d(x, Σn) dμ(x)

R dWn, μ)→0 asn→ ∞. We note that the bound does not depend onΣn. Therefore, using Lemma1,

lim inf

n→∞ Eμnn)=lim inf

n→∞ Eμn)Eμ(Σ ). 2

Corollary 4.Assume thatμn μ inPR and thatΣn is a minimizer ofEμn. Then along a subsequenceΣndH Σ whereΣis a minimizer ofEμ.

Proof. Since by Blaschke’s theorem (see[9]) the sequenceΣn has a subsequence which converges in Hausdorff metric the claim follows from theΓ convergence. 2

Corollary 5.AssumeEμhas a unique minimizerΣand thatμn μ inPR. Then for everyε >0there existsn0such that for allnn0anyminimizerΣnofEμn satisfiesdH(Σ, Σn) < ε.

Proof. Assume that the claim does not hold. Than there existsε >0 and a sequenceΣnk of minimizers of Eμnk such that for eachk,dH(Σ, Σnk)ε. By relabeling, we can assumenk=kfor allk. By Blaschke’s theorem, there existsΣ˜ ∈Asuch that, along a subsequence,Σn→ ˜Σasn→ ∞in Hausdorff metric. We can again assume that the subsequence is the whole sequence. FurthermoreΣ˜ is connected and thus belongs toA. We note thatdH(Σ,Σ)˜ ε.

By the lower-semicontinuity part in theΓ-convergence,Σ˜ is a minimizer of Eμ, which contradicts the uniqueness assumption. 2

(5)

Lemma 6.LetR >0. Letγn: [0,1] →B(0, R)be a sequence of Lipschitz curves with constant-speed parameteriza- tion(i.e.|γn(s)| =length(γn)for a.e.s∈ [0,1]). Assume thatsupnlength(γn)andsupnγnBV are finite. Then along a subsequenceγnconverges to a Lipschitz curveγ in the sense that

γnγ inCα asn→ ∞, for anyα∈ [0,1), γnγ inLpasn→ ∞, for anyp∈ [1,∞), and

γn γ in the space of finite signed Borel measures asn→ ∞.

Proof. The constant-speed assumption and the uniform bound on the lengths imply that γnL are uniformly bounded and thus there is a uniform bound on the Lipschitz norm for the curves. The fact thatγn converges along a subsequence inCαfor anyα∈ [0,1)follows since the set of Lipschitz functions with values inB(0, R), is compactly embedded in Cα. To obtain the convergence that holds for allα at the same time one also uses a diagonalization argument.

From the embedding theorem of BV spaces (see [9,10]), it follows that for some gBV ([0,1],Rd), along a further subsequence,γng inL1as n→ ∞andγn g in the space of signed measures asn→ ∞. Using the definition of the weak derivative it follows thatg=γ. SinceγnLare uniformly bound by interpolation it follows that for allp∈ [1,∞),γnγinLpasn→ ∞. Furthermore|γn| → |γ|inL1asn→ ∞and the constant-speed assumption imply that and moreover|γ(s)| =length(γ )for a.e.s∈ [0,1]and in particularγ is a Lipschitz curve. 2 2.1. Scaling ofEμwith respect to dilations ofμ

Given a setA⊂Rdandr >0 we defineAr = {x: rxA}.

Given a measureμ andr >0 we define the dilation of μto scale r to be the measureDrμ such that for any μ-measurable setA

Drμ(A)=μ A

r .

We note that since both terms ofEare scale linearly with respect to length Eμ(Σ )=1

rEDrμ(rΣ ).

Therefore ifΣis a minimizer ofEμthenis a minimizer ofEDrμ. 3. Discrete data

In this section we consider the case thatμis a discrete (or particle) measure:

μ= n

i=1

miδxi (5)

wheremi>0 andxi∈Rd. The measuresmiδxi are calledparticles. We denote the support ofμbyX= {x1, . . . , xn}.

Lemma 7.Ifμis discrete then every minimizerΣis graph with straight edges.

Proof. Letϕ:XΣbe the mapping that assigns to eachxiXa point onΣwhich is the closest toxi(if the closest point is nonunique, an arbitrary one is chosen). Letyi=ϕ(xi)andY = {y1, . . . , yn}(the points are not necessarily distinct). LetAbe the Steiner tree containing the setY. TheSteiner treeis the connected graph with minimal total length of edges containing the vertices inY (it can have other vertices as well). We note that it is also the connected set of minimalH1measure containingY. For further information on Steiner trees we refer to[11]and[12].

Furthermore note thatE(A)E(Σ )with equality holding only ifΣis also a Steiner tree, which proves the claim.

We remark thatΣmay be different thanAsince Steiner trees are not necessarily unique. 2

(6)

Now that we know thatΣ has straight edges, we can study it more carefully. We define the vertices,V as the set of those pointsvinΣfor which there exists a pointx inXsuch thatvis the closest point toxinΣ:

V =

vΣ: (xX) (zΣ ) d(x, v)d(x, z)

. (6)

Note thatYV and that it is possible that at a vertex of degree two the angle is 180. Since, by above the segments of Σ connecting the vertices must be line segments, we define edges,S, as follows: forv, wV,{v, w} ∈S is an edge if the line segment[v, w] ⊆Σ. We note that sinceΣis made of finitely many line segments,V must be finite.

Thus we can writeV = {v1, . . . , vm}. Since

Rd

d(x, Σ ) dμ(x)= n

i=1

mid(xi, Σ )= n

i=1

mid(xi, V ),

Σmust be connected the graph of minimal total length containing the verticesV. That isΣ is a Steiner tree[11]for the setV too.

The following facts on Steiner trees are available in the classical paper by Gilbert and Pollak[11].

Proposition 8.LetG=(V , S)be as above. Then (i) Gis a tree, that is it does not contain a closed loop.

(ii) If{u, v}and{v, w}are edges then the angle uvw120. (iii) The maximal degree of a vertex is three.

(iv) Ifvis a vertex of degree three then the angles between edges atvare120, and thus all three edges belong to a 2-dimensional plane.

We call the vertices of degree one theendpoints, the ones of degree twocorners, and the ones of degree threetriple junctions. Givenj=1, . . . , mletIj be the set of indices of points inXfor whichvj is the closest point inV

Ij=

i∈ {1, . . . , n}: (k=1, . . . m) d(xi, vj)d(xi, vk)

=

i∈ {1, . . . , n}: (yΣ ) d(xi, vj)d(xi, y)

. (7)

IfiIj then we say thatxi talks tovj. We say that a vertexvj istied downif for somei,vj=xi. We then say that vjistied toxi. Note that ifvj is tied toxitheniIj andTij=mi. A vertex which is not tied down is calledfree. We show below that ifxi talks tovj andvj is free thenxicannot talk to any other vertex.

Consider annbymmatrixT such that

Tij0, m

j=1

Tij=mi, and Tij>0 implies iIj. (8)

Note thatμ=n i=1

m

j=1Tijδxi. We defineσ to be the projection ofμonto the setΣ, in the sense that the mass fromμis transported to a closest point onΣ. That is

σ= m

j=1

n

i=1

Tijδvj. (9)

We note that the matrix T describes an optimal transportation plan betweenμ and σ with respect to any of the transportation costsc(x, y)= |xy|p, for p1. We claim that such matrix T exists. It is enough to consider mappingϕfrom the proof of Lemma7and then define

Tij=mi ifϕ(xi)=vj, 0 otherwise.

(7)

We note that in this discrete setting E(Σ )=

n

i=1

mid(xi, Σ )+λ

{v,w}∈S

|vw|

= m

j=1

iIj

Tij|xivj| +λ

{vj,vk}∈S

|vjvk|. (10)

Lemma 9.Assume thatΣminimizesEfor discreteμdefined in(5). LetV be the set of vertices as defined in(7)and T be any matrix(transportation plan)satisfying(8). For any vertexvj:

(i) Ifvj is an endpoint then letwbe the vertex such that{vj, w}is an edge. Ifvj is free then

iIj

Tij

xivj

|xivj|+λ wvj

|wvj| =0. (11)

Ifvj is tied toxkthen

iIj,i=k

Tij

xivj

|xivj|+λ wvj

|wvj|

mk. (12)

(ii) Ifvj is a corner then let{w1, vj}and{vj, w2}be the edges at the corner. Ifvj is free then

iIj

Tij xivj

|xivj|+λ

w1vj

|w1vj|+ w2vj

|w2vj| =0. (13)

Ifvj is tied toxkthen

iIj,i=k

Tij xivj

|xivj|+λ

w1vj

|w1vj|+ w2vj

|w2vj| mk. (14)

(iii) Ifvj is a triple junction and ifvj is free then

iIj

Tij xivj

|xivj|=0. (15)

Ifvj is tied toxkthen

iIj

Tij xivj

|xivj|

mk. (16)

Proof. To prove (i) and (ii), consider now the configuration Σv which is obtained from Σ just by changing the location ofvjtov. LetSvbe the set of edges of the new graph. Formulation (10) provides

E(Σv)

iIj

Tij|xiv| +λ

{v,vl}∈Sv

|vvl| + m

l=1,l=j

iIl

Til|xivl| +λ

{vk,vl}∈Sv

|vkvl|

=:F (v).

Note thatF defined above mapsRd→Rand thatE(Σ )=F (vj).

Ifvj is a free vertex thenF is a smooth function nearvj. The minimality ofΣ thus implies that DF (vj)=0.

Straightforward computation ofDF gives the conditions (11) and (13).

If on the other handvj is tied toxkfor somek,xk=vj then recall thatkIj and furthermoreTkj =mk.F is no longer smooth atvj but it still has a minimum atvj. Therefore zero vector must belong to the subgradient ofF atvj,

(8)

that is 0∈∂F (vj). Using that the subgradient at 0 ofz→ |z|isB(0,1)we conclude that 0∈∂F (vj)if the conditions (12) and (14) hold at an endpoint and corner, respectively. More precisely if we define

F (v)˜ =

iIj, i=k

Tij|xiv| +λ

{v,vl}∈Sv

|vvl|. Then, usingvj=xk,

F (v)=mk|vjv| + ˜F (v)+const.

and hence

∂F (vj)=B(0, mk)+DF (v˜ j).

Therefore 0∈∂F (vj)if|DF (v˜ j)|mk.

Obtaining (15) and (16) is analogous, only that one also needs the fact that the angles at triple junction are 120, see Proposition8. 2

Corollary 10.Assume that conditions of the lemma are satisfied.

(i) In two dimensions,d=2, ifvj is a triple junction and is a free vertex thenIj= ∅.

(ii) Ifitalks tovj andvk (j=k)then bothvj andvk are tied down. Consequently ifxi talks tovj andvj is free thenTij=mi. Furthermoremn.

(iii) Ifvjis an endpoint then

iIj

Tijλ. (17)

Hence at every endpoint, the measureσ defined in(9)has an atom of the mass at least equal toλ. Note that this gives an upper bound on the number of endpoints. A further consequence is that if2λ >n

i=1mi then the minimizerΣis just a single point(which is then a vertex of degree zero).

Proof. To prove (i) assume thativj. ThenB(xi,|vjxi|)Σ= ∅. But this contradicts the fact that the angles at triple junction are 120.

To prove (ii) assume that there exist i∈ {1, . . . , n}andj, k distinct elements of {1, . . . , m}such thatiIj and iIk. LetT be a matrix satisfying the condition (8). For s(0,1)consider the matrixT (s)obtained from T by setting:

Tij(s)=mi(1s), Tik=mis and Til=0 ifl /∈ {j, k}. Note thatT (s)satisfies the condition (8).

We argue by contradiction: assume thatvjis free. Let us consider first the case thatvj is an endpoint. Letwbe the vertex for which{v, w}is an edge. The condition (11) must be satisfied forT (s)for alls∈ [0,1]:

mi(1s) xivj

|xivj|+

lIj,l=i

Tlj

xlvj

|xlvj|+λ wvj

|wvj|=0.

One arrives at a contradiction by taking the derivative ins.

The proofs for a corner and for a triple junction are analogous.

Let us explain why the above impliesmn. By definition ofV, for eachvjV,Ij= ∅, that there is a particle talking tovj. We claim that for everyvj there is a particle talkingonlytovj. Ifvj is free then by above this is the case with any particle inIj. Ifvj is tied toxkthenxkonly talks tovj. Consequently there must be more particles then vertices inV.

Let us now consider the claim of (iii). There are two cases. We first consider the case thatvj is free. Then (11) holds. Taking the inner product with |wwvvj

j| and using (ii) above gives

iIj

mi xivj

|xivjwvj

|wvj|=λ.

(9)

Thus

iIj

mi

iIj

mi xivj

|xivjwvj

|wvj| λ.

Ifvj is tied toxkfor somekthen the condition (12) gives

iIj,i=k

Tij

xivj

|xivjwvj

|wvj|+λ mk. Thus

λmk

iIj,i=k

Tij xivj

|xivjwvj

|wvj|

iIj

Tij. 2

3.1. Turning angle

Ifv is a corner with adjacent verticesw1andw2then theturning angleatvisTA(v)=π w1vw2. Basically it describes the curvature atv. ForAΣ, the turning angle ofA,TA(A)=

vAVTA(v). The total turning angle, TTA=TA(Σ ).

Lemma 11.IfΣis a minimizer ofEandAΣ.

TA(A) π

i∈∪{Ij:vjA}

mi. (18)

Proof. Let us first consider the case that Ais a single corner, A= {vj}. Let w1 andw2be the adjacent vertices.

Letαbe the half of the turning angle:TA(vj)=2α. Then w1vjw2=π−2α. Letθ1=|vvjjww11| andθ2=|ww22vvjj|. Elementary geometry yields that|θ2θ1| =2 sinα.

Analogously to the proof of statement (iii) of Corollary10, that is by using conditions (13) and (14) and taking inner product with |θθ2θ1

2θ1|, one can show that 2λsinα

iIj

Tij. (19)

Therefore αmax

π 2,arcsin

iIjTij

π

2

iIjTij 2λ which establishes the claim.

For the generalAΣit suffices to sum over the vertices it contains. 2 3.2. Region of influence

Given orthogonal unit vectorsξandband an angleβ∈ [0,π2], consider the wedgeW, with bisectorband opening 2β, defined as follows:

W (ξ, b, β)=

x∈Rd: |ξ·x|< b·xtanβ

. (20)

ByJz1, . . . , zkKwe denote the piecewise linear curve connecting the pointsz1, . . . , zk. Given three pointsv1,v2, andv3 such that if they are collinear thenv2 lies between v1 andv3 considerΣ=Jv1, v2, v3K. If the points are collinear the region of influence ofv2, defined in (3), is the hyperplane passing throughv2, orthogonal tov3v2. If points are not collinear then letθi =|vvii+1+1vvi

i| for i=1,2. The region of influence ofv2 the translated wedge:

R(v2)=v2+W (ξ, b, β)whereξ =|θθ11++θθ22|,b=|θθ11θθ22| andβ=TA(v2)/2. We denote the mapping above that outputs

(10)

Fig. 1. The basic configuration.

the Normal,Bisectris, andAngle given non-collinear pointsv1, v2, v3 by(ξ, b, β)=NBA(v1, v2, v3). Then we can writeR(v2)=v2+W (NBA(v1, v2, v3)).

We note that ifΣ=Jv1, . . . , vmKthen for alli=2, . . . , m−1, ifTA(vi) >0R(vi)vi+W (NBA(vi1, vi, vi+1)).

3.3. Basic configuration

Here we describe the basic configurationμ, whose perturbation is used in the counterexample to regularity. While¯ the construction is rather flexible we choose a fixed configuration to make the number of parameters of the system easier to manage.

Letm1=m3=0.38,m2=0.24, andλ=0.36. Letx¯1=(−1,0),x¯2=(0,1), andx¯3=(1,0),

¯

μ=m1δx¯1+m2δx¯2+m3δx¯3. (21)

Note that the total mass is one.

We now turn to characterizing the minimizer. We note that sinceλis greater than one third of the total mass the condition (17) implies that the minimizer can have at most two endpoints. Thus the minimizer is either a point or a piecewise linear curve. We reindex the verticesv1, . . . , vmif needed so that the minimizer isJv1, . . . , vmK.

Let us note that for any pointa∈R2,E({a}) > E(Jx¯1, a,x¯3K)and thus a point cannot be a minimizer. Therefore a minimizer is a piecewise linear curve with either one or two line segments. If it has two line segments thenv2= ¯x1 since then it would violate angle condition (ii) of Proposition 8, since we know the minimizer must stay in the convex hull of the support ofμ. If¯ v1= ¯x1andvm= ¯x1thenE(Jx¯1, v1, . . . , vmK) < E(Jv1, . . . , vmK)sincem1> λ.

Analogously forx¯3. Thereforex¯1 andx¯3must be endpoints of any minimizer. So without loss of generality we can setv1= ¯x1andvm= ¯x3. IfJv1, v2Kwere a minimizer thenx¯2would talk to(0,0)so(0,0)would have to be a vertex of the graph (as we defined it (6)). But then the condition (13) cannot hold. So the minimizer must be of the form [ ¯x1,v¯2,x¯3].

The criticality condition (13) implies that the only minimizer is the symmetric configuration,Σ, presented on Fig.1. Elementary geometry gives:

¯

v2=(0, H ), whereH= 1 2√

2, L= 3

2√

2, and sinα=1

3. (22)

4. Construction of the counterexample

We start with the basic configurationμ¯ defined in (21), which we now consider as configuration inRdby taking the values of coordinates from 3 tod to be zero.

To smooth out the basic configuration, we use a standard mollifierη. That is, letηbe smooth, radially symmetric, positive onB(0,1), equal to zero outside ofB(0,1), and such that

Rdη(x) dx=1. Forδ >0 letηδ(z)=δ1dη(zδ). Let ρi,δ=miηδ(· − ¯xi)fori=1,2,3 and letμδbe the measure with densityρ1,δ+ρ2,δ+ρ3,δ.

To have a measure with connected and convex support we introduce the background measureμ˜ to be the measure with densityη1.5. The smooth measure we consider is

μq,δ=(1q)μδ+qμ.˜

Theorem 12.There existδ >0and1> q >0such that there is a minimizer ofEforλ=0.36andμ=μq,δwhich is a Lipschitz curve such that if one considers its constant-speed parameterizationγ: [0,1] →Rd(|γ(s)| =length(γ )

(11)

for a.e.s∈ [0,1])thenγis anRdvalued BV function such thatγhas an atom of size at least1at somes(0,1).

More precisely|γ({s})|1.

Proof. Assume thatεsatisfies the condition (C1) 0< ε <2000α .

Corollary 5 implies that forδ >0 andq >0 small enough any minimizer ofEμq,δ lies within ε ball of Σ in Hausdorff metric. That is, we can impose

(C2) q >0 andδ >0 are small enough so that any minimizerΣq,δofEμq,δ satisfiesdHq,δ, Σ) <ε2. We also require:

(C3) q <π 20000α . (C4) δ <0.1ε.

The condition (C3) controls the part of the turning angle which is due to the background measure.

Step1o. Discrete approximation.Letμq,δ,n be an approximation ofμq,δ which is a particle measure such that the Wasserstein distancedWq,δ, μq,δ,n) <1n and furthermore that there exists an optimal coupling such that all of the mass in the(1q)ρi,δpart ofμq,δis coupled with particles inB(x¯i, δ)fori=1,2,3. This can be achieved by, say, taking a fine square grid such thatx¯1,x¯2, andx¯3are cell centers and then constructingμq,δ,n by taking the mass of μq,δin each grid cell and concentrating it at the center of the grid cell.

Due to Corollary4, along a subsequence, minimizers,Σq,δ,n (if nonunique then any minimizer can be chosen), converge in Hausdorff metric to a minimizerΣq,δofEμq,δ. We can assume without loss of generality that the whole sequence converges and thatnis large enough so that

dHq,δ,n, Σq,δ) < ε 2

which implies, using (C2), thatdHq,δ,n, Σ) < ε.

Step2o. Control of the optimal coupling.We claim that particles ofμq,δ,n which lie inB(x¯2, δ)can only talk to points onΣq,δ,nwhich lie above (in the direction ofe2) the hyperplaneP = {y: y·e2=Hδcosεα}.

To prove the claim considerxiB(x¯2, δ). Letx˜i be the projection ofxi on the coordinate axis in the direction of the vectore2. That isx˜i=(0, xi,2,0, . . . ,0). LetU= {z: d(z, Σ) < εandz·e1=0}. We note thatUΣq,δ,n= ∅. An elementary geometry argument shows that the furthest point tox˜2onUis(0, Hcosεα,0, . . . ,0). Thus

d(xi, Σq,δ,n)d(xi,x˜i)+d(x˜i, Σq,δ,n) < δ+xi,2H+ ε cosα. On the other hand

d(xi, P )=xi,2+δH+ ε cosα.

Step3o. Average tangent direction.SinceΣq,δ,nhas only two endpoints, and is piecewise linear, it can be repre- sented by a constant-speed parameterized curveγq,δ,n: [0,1] →Rd. Due to the closeness toΣ(by Step 1) there are points onΣq,δ,nwhich are withinεofx¯1andx¯3.

We claim that the endpoints of the curve must lie within 2εofx¯1andx¯3. For if that was not the case then all of the mass in, say,B(x¯1, δ)would not talk to an endpoint. But then there would not be enough available mass for the lower bound on the mass talking to the endpoints (condition (17)) to be satisfied.

We can require thatγq,δ,n(0)lies close tox¯1. We define θ (s)= γq,δ,n (s)

|γq,δ,n (s)| = γq,δ,n (s) length(γq,δ,n).

(12)

Fig. 2. Schematic illustration of minimizers ofEμ¯,Eμq,δ, andEμq,δ,n. Some lengths are distorted to achieve better clarity of the illustration.

We note thatθBV ([0,1],Rd). Let θ¯1= v¯2− ¯x1

v2− ¯x1| and θ¯2= x¯3− ¯v2

| ¯x3− ¯v2|.

Let Cyl1= {x ∈Rd: d(x, Σ) < εand(x− ¯x1)· ¯θ1∈ [2ε, L−11ε]} and Cyl2= {x ∈Rd: d(x, Σ) < ε and

(x− ¯v2)· ¯θ2∈ [2ε, L−11ε]}, as shown on Fig.2. Lets1,1,n be the first timeγq,δ,n enters Cyl1 ands1,2.n the largest timeγq,δ,n belongs to Cyl1. Analogouslys2,1,n ands2,2,n are the corresponding times for the domain Cyl2. Letθi,avg=|γγq,δ,nq,δ,n(s(si,2,ni,2,n))γγq,δ,nq,δ,n(s(si,1,ni,1,n))| fori=1,2. We claim that θ¯iθi,avg<0.01αfori=1,2. To see this, note that tan θ¯iθi,avg is less than twice the width of the cylinder Cyli divided by its length: tan θ¯iθi,avg< L13ε <L <

0.005αby (C1), which implies the claim.

Step4o. Tangent at the end of the cylinders.The cylinders Cyl1and Cyl2were defined so that they lie below the hyperplaneP, as can be verified by simple trigonometry. Step 2 then implies that no point that belongs toB(x¯i, δ) for i=1,2,3 talks to any point in the cylinders. Thus, using the turning angle estimate (18) and assumption (C3), TA(CyliΣq,δ,n)π q <100α . Therefore the tangent at the any point in Cyliis close to the average tangent. Let

θ1=γq,δ,n (s1,2,n+)

length(γq,δ,n) := lim

ss1,2,n

γq,δ,n (s)

length(γq,δ,n) and θ2=γq,δ,n (s2,1,n) length(γq,δ,n)

be the tangents as the curve exits Cyl1and as it enters Cyl2. Then θiθi,avg<0.01α. Combining with estimates of Step 3 we get

¯θiθi ¯θiθi,avg+ θi,avgθi<0.02α fori=1,2.

Step5o. The turning angle at the first contact point.We now show that there exists a vertex at which the turning angle is large (of size at least aboutα/2). This is the key point of the argument. Let us relabel the vertices if necessary, so that their indices are increasing alongγq,δ,n ass increases. Letvj be the first vertex of γq,δ,n that talks to any particle ofμq,δ,n inB(x¯2, δ), as illustrated on Fig.3. The reason that the turning angle has to be “large” is that the tangent cannot turn much prior tovj (because the vertices talk to very little mass), but forvj to be able to talk to a point inB(x¯2, δ)the tangent must turn by at least by aboutα. Thus it has to turn by that amount precisely atvj.

Here is the detailed argument. Let θ= vjvj1

|vjvj1| and θ+= vj+1vj

|vj+1vj|.

Letsnbe the time at whichγq,δ,n(sn)=vj. Since points onγq,δ,n|[s1,2,n,sn) can only talk to the background measure qμ, it follows from (18) that˜ TA(γq,δ,n([s1,2,n, sn))) <π q <0.01α. Combining with Step 4 implies

¯θ1θ<0.03α. (23)

LetK= {x: d(x, Σ) < ε, x·e2> Hδcosεα}. From Step 2 follows thatvjK. LetyB(x¯2, δ). We can decompose vectors inRdinto the component in the direction ofe2(vertical component) and the one in the orthogonal complement ofe2(horizontal component). Elementary geometry implies using the assumptions (C1) and (C4) that

Références

Documents relatifs

We will study algebraic hyper-surfaces on the real unit sphere S n − 1 given by an homogeneous polynomial of degree d in n variables with the view point, rarely exploited, of

and for interval distances [10], the pruning is done using only the known distances of a current node to the previous fixed nodes, with our strategy we aim to prune the search tree

GIUSTI, Local estimates for the gradient of non-parametric surfaces of prescribed mean curvature, Comm. GERHARDT, Existence, regularity, and boundary behaviour of

any 11-subspace of L’(v) which is generated by a sequence of prob- abilistically independent random variables is complemented?. A slight improvement of this result will be

Nevertheless, to construct a counter-example to Cantelli conjecture, we will consider probability measures such that a bounded Brownian transport exists (and the proof of its

Liberti, L., Lavor, C., Mucherino, A.: The discretizable molecular distance geometry problem

We use this mv-decomposition to propose a solution to the distance-2 broadcast problem in a synchronous multi-hops radio networks where adjacent transmissions are subject

We suppose that we know that the solutions of the planar isoperimetric problem are circles. We consider a very old variant of this problem called Dido’s problem [18]. It is related