• Keine Ergebnisse gefunden

Γ-Limits of convolution functionals

N/A
N/A
Protected

Academic year: 2021

Aktie "Γ-Limits of convolution functionals"

Copied!
31
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Γ-limits of convolution functionals

Luca Lussardi and Annibale Magni

Preprint 2010-11 November 2010

Fakult¨ at f¨ ur Mathematik

Technische Universit¨ at Dortmund Vogelpothsweg 87

44227 Dortmund tu-dortmund.de/MathPreprints

(2)
(3)

Γ-LIMITS OF CONVOLUTION FUNCTIONALS

LUCA LUSSARDI AND ANNIBALE MAGNI

Abstract. We compute the Γ-limit of a sequence of non-local integral functionals de- pending on a regularization of the gradient term by means of a convolution kernel. In particular, as Γ-limit, we obtain free discontinuity functionals with linear growth and with anisotropic surface energy density.

Keywords: Free discontinuities, Γ-convergence, anisotropy.

2010 Mathematics Subject Classification: 49Q20, 49J45, 49M30.

Contents

1. Introduction 1

2. Notation and preliminaries 4

2.1. Functions of bounded variation 4

2.2. Slicing 5

2.3. Γ-convergence 5

2.4. Supremum of measures 6

2.5. A density result 6

2.6. A relaxation result 6

3. Statement of the main results 7

3.1. The anisotropy 8

3.2. Main results 9

4. Compactness 10

5. The Γ-liminf inequality 11

5.1. A preliminary estimate from below in terms of the volume and Cantor parts 11

5.2. A preliminary estimate in terms of the surface part 15

5.3. Proof of the Γ-liminf inequality 19

6. The Γ-limsup inequality 21

7. Computation of θin the one-dimensional case 24

References 26

1. Introduction

As it is well known, many variational problems which are recently under consideration, arising for instance from image segmentation, signal reconstruction, fracture mechanics and liquid crystals, involve afree discontinuity set (according to a terminology introduced in [19]). This means that the variable functionuis required to be smooth outside a surfaceK, depending onu, and bothu andK enter the structure of the functional, which takes the form given by

F(u, K) = Z

Ω\K

φ(|∇u|) dx+ Z

K∩Ω

θ(|u+−u|, νK) dHn−1,

being Ω an open subset ofRn, K is a (n−1)-dimensional compact subset ofRn, |u+−u| the jump ofuacrossK,νKthe normal direction toK, whileφandθgiven positive functions, whereas Hn−1 denotes then−1-dimensional Hausdorff measure.

1

(4)

The classical weak formulation for such problems can be obtained consideringK as the set of the discontinuities ofuand thus working in the space of functions with bounded variation. More precisely, the aforementioned weak form ofF takes onBV(Ω) the general form

(1.1) F(u) =

Z

φ(|∇u|) dx+ Z

Su

θ(|u+−u|, νu) dHn−1+c0|Dcu|(Ω),

where Du =∇uLn+ (u+−u)Hn−1+Dcuis the decomposition of the measure derivative of u in its absolutely continuous, jump and Cantor part, respectively, and Su denotes the set of discontinuity points ofu.

The main difficulty in the actual minimization ofF comes from the surface integral Z

Su

θ(|u+−u|, νu) dHn−1,

which makes it necessary to use suitable approximations guaranteeing the convergence of minimum points and naturally leads to Γ-convergence.

As pointed out in [10], it is not possible to obtain a variational approximation for F by the typical integral functionals

Fε(u) = Z

fε(∇u) dx

defined on some Sobolev spaces. Indeed, when considering the lower semicontinuous envelopes of these functionals, we would be lead to a convex limit, which conflicts with the non-convexity of F.

Heuristic arguments suggest that, to get rid of the difficulty, we have to prevent that the effect oflargegradients is concentrated onsmall regions. Several approximation methods fit this requirements. For instance in [7], [12], [24] the case where the functionals Fε are restricted to finite elements spaces on regular triangulations of sizeεis considered. In [1], [2], [23] the implicit constraint on the gradient through the addition of a higher order penalization is investigated.

Moreover, it is important to mention theAmbrosio & Tortorelliapproximation (see [4] and [5]) of the Mumford-Shah functional via elliptic functionals.

The study of non-local models, where the effect of a large gradient is spread onto a set of size ε, was first introduced by Braides & Dal Maso in order to approximate the Mumford-Shah functional (see [10] and also [11], [13], [14], [15], [16]) by means of the family

(1.2) Fε(u) = 1

ε Z

f

ε Z

Bε(x)∩Ω

|∇u|2dy

dx, u∈H1(Ω),

where, for instance,f(t) =t∧1/2 and Bε(x) denotes the ball of centrexand radiusε. A variant of the method proposed in [10] has been used in [22] to deal with the approximation of a functional F of the form (1.1), with φhaving linear growth and θ independent on the normal νu (see also [20] and [21]). More precisely, in [22] the Γ-limit of the family

Fε(u) =1 ε

Z

f

ε Z

Bε(x)∩Ω

|∇u|dy

dx, u∈W1,1(Ω), for a suitable concave functionf, is computed.

In [25] (see also [13]) the case of an anisotropic variant of (1.2) has been considered. In particular it is proven that the family

Fε(u) = 1 ε Z

f ε|∇u|p∗ρε

dx, u∈H1(Ω), p >1, Γ-converges to an anisotropic version of the Mumford-Shah functional.

In this paper we investigate the Γ-convergence of the family Fε(u) = 1

ε Z

fε ε|∇u| ∗ρε

dx, u∈W1,1(Ω).

The main difficulty to overcome is the estimate from below for the lower Γ-limit in terms of the surface part, while the contribution arising from the volume and Cantor parts has been treated along the same line of the argument already exploited in [25]. The estimate from above has

(5)

been achieved by density and relaxation arguments. We prove that the Γ-limit, in the strong L1-topology, is given by

F(u) = Z

φ(|∇u|) dx+ Z

Su

θ(|u+−u|, νu) dHn−1+c0|Dcu|(Ω), wherec0= limt→+∞φ(t)/tand

θ(s, ν) = inf

lim inf

j→+∞

1 εj

Z

Qν

f(εj|∇uj| ∗ρεj) dx: (uj)∈Wν0,s, εj→0+

,

being Wνa,b the space of all sequences on the cylinder Qν which converge, shrinking onto the interface, to the function that jumps fromatobaround the origin (see paragraph 3.1 for details).

In section 7 we have been able to show that the method used in [22] to writeθ in a more explicit form works only if n = 1. In the case n > 1 such an argument does not work. Let us briefly discuss the reason. Without loss of generality we can supposeν =e1. LetPC be the orthogonal projection of Conto{x1 = 0}. Denote byX the space of all functions v∈Wloc1,1(R×PC) which are non-decreasing in the first variable and such that there existξ0< ξ1with v(x) = 0 ifx1< ξ0

andv(x) =sifx1> ξ1. Then, exploiting the same argument as in [22], we haveθ(s,e1)≥infXG, where

G(v) = Z +∞

−∞

f Z

C(se1)

1v(z)ρ(z−te1) dz

dt.

The estimateθ(s,e1)≥infXGturns out to be optimal if infXG= infY G, whereY is the space of all functions v ∈ X such that v depends only on the first variable. This is due to the fact that proving the inequality θ(s,e1) ≥ infXG we lose control on all the derivatives ∂iv for any i = 2,· · ·, n. In the case C = B1 and ρ = ω1

n

χB

1, treated in [22], one is able to prove that infXG= infY G computing directly infXG by a discretization argument (see Prop. 5.7 in [22]).

In general, infXG= infYGdoes not hold. Indeed proceeding at first as in the proof of Prop. 5.6 in [22], one is able to show that for any C ⊂ R2 open, bounded, convex and symmetrical set (i.e.C=−C) and forρ= |C|1 χC, it holds

(1.3) inf

Y G= Z h1

−h1

f s

|C|H1(C∩ {z1=t}

dt.

Now ifC is the parallelogramC={(x, y)∈R2:−2≤y≤2, x−1≤y≤x+ 1}applying (1.3), we get

infY G0 = 2f 2s

|C|

+ 2

Z 2 0

f sr

|C|

dr.

If we computeGon the functionwgiven by w(x, y) =

0 ify > x−1 s ify≤x−1 ,

(to do this we notice that the functionalG makes sense also onBVloc(R×(−2,2)) writing D1v instead of∂1vdz) we obtain

G(w) = 2f 4s

|C|

.

Iff is strictly concave then G(w)<2f

2s

|C|

+ 2f

2s

|C|

<2f 2s

|C|

+ 2

Z 2 0

f sr

|C|

dr= inf

Y G.

By a density argument we deduce that infXG <infY G.

As a conclusion, it seems that for a generic anisotropic convolution kernelρεthe expression for θcan not be further simplified whenn >1.

(6)

2. Notation and preliminaries

We will denote byLp(Ω) and byWk,p(Ω), fork∈N,k≥1, and for 1≤p≤+∞, respectively the classical Lebesgue and Sobolev spaces on Ω. The Lebesgue measure of a measurable setA⊂Rn will be denoted by|A|, whereas the Hausdorff measure ofAof dimensionm < nwill be denoted by Hm(A). The ball centered inxwith radius r will be denoted by Br(x), while Br stands for Br(0); moreover, we will use the notationSn−1for the boundary ofB1 inRn. The volume of the unit ball inRn will be denoted byωn, with the convention ω0= 1. FinallyA(Ω) denotes the set of all open subsets of Ω.

2.1. Functions of bounded variation. For a thorough treatment ofBV functions we refer the reader to [3]. Let Ω be an open subset of Rn. We recall that the spaceBV(Ω) of real functions of bounded variation is the space of the functions u ∈ L1(Ω) whose distributional derivative is representable by a measure in Ω,i.e.

Z

u∂ϕ

∂xi

dx=− Z

ϕdDiu, ∀ϕ∈Cc(Ω),∀i= 1, . . . , n,

for someRn-valued measureDu= (D1u, . . . , Dnu) on Ω. We say thatuhasapproximate limitat x∈Ω if there existsz∈Rsuch that

lim

r→0+

Z

Br(x)

|u(y)−z|dy= 0.

The setSu where this property fails is called approximate discontinuity setof u.The vectorz is uniquely determined for any point x∈Ω\Su and is called theapproximate limit of uat xand denoted by ˜u(x).We say thatxis anapproximate jump pointof the functionu∈BV(Ω) if there exista, b∈Randν ∈Sn−1 such thata6=band

(2.1) lim

r→0+

Z

B+r(x,ν)

|u(y)−a|dy= 0, lim

r→0+

Z

Br(x,ν)

|u(y)−b|dy= 0,

whereBr+(x, ν) ={y∈Br(x) : hy−x, νi>0}andBr(x, ν) ={y∈Br(x) : hy−x, νi<0}.The set of approximate jump points ofuis denoted by Ju.The triplet (a, b, ν),which turns out to be uniquely determined up to a permutation ofaandb and a change of sign ofν,is usually denoted by (u+(x), u(x), νu(x)). On Ω\Su we set u+ =u = ˜u.It turns out that for any u∈BV(Ω) the setSu is countably (n−1)-rectifiable andHn−1(Su\Ju) = 0.Moreover,

Du Ju= (u+−uuHn−1 Ju

andνu(x) gives the approximate normal direction toSuforHn−1-a.e. x∈Su.

For a function u∈ BV(Ω) let Du = Dau+Dsu be the Lebesgue decomposition of Du into absolutely continuous and singular part. We denote by ∇u the density of Dau; the measures Dju:=Dsu JuandDcu:=Dsu (Ω\Su) are called thejump partand theCantor partof the derivative, respectively. It holdsDu=∇uLn+ (u+−uuHn−1 Ju+Dcu. Let us recall the following important compactness Theorem inBV (see Th. 3.23 and Prop. 3.21 in [3]):

Theorem 2.1. Let Ω be a bounded open subset of Rn with Lipschitz boundary. Every sequence (uh)inBV(Ω)which is bounded inBV(Ω)admits a subsequence converging inL1(Ω)to a function u∈BV(Ω).

We say that a function u ∈ BV(Ω) is a special function of bounded variation, and we write u ∈ SBV(Ω), if |Dcu|(Ω) = 0. We say that a function u ∈ L1(Ω) is a generalized function of bounded variation, and we write u∈GBV(Ω), ifuT := (−T)∨u∧T belongs toBV(Ω) for every T ≥0. Ifu∈GBV(Ω), the function∇ugiven by

(2.2) ∇u=∇uT a.e. on{|u| ≤T}

turns out to be well-defined. Moreover, the set functionT 7→SuT is monotone increasing; therefore, if we set Su = S

T >0JuT , for Hn−1-a.e.x ∈ Su we can consider the functions of T given by (uT)(x), (uT)+(x),νuT(x). It turns out that

(2.3) u(x) = lim

T→+∞(uT)(x), u+(x) = lim

T→+∞(uT)+(x), νu(x) = lim

T→+∞νuT(x)

(7)

are well-defined for Hn−1-a.e.x ∈ Su Finally, for a function u ∈ GBV(Ω), let |Dcu| be the supremum, in the sense of measures, of |DcuT| for T >0. It can be proved that for any Borel subsetB of Ω

(2.4) |Dcu|(B) = lim

T→+∞|DcuT|(B).

2.2. Slicing. In order to obtain the estimate from below of the lower Γ-limit (see next paragraph) we need some basic properties of one-dimensional sections of BV-functions. We first introduce some notation. Letξ ∈Sn−1, and let ξ be the vector subspace orthogonal to ξ. If y ∈ξ and E⊆Rnwe setEξ,y={t∈R: y+tξ∈E}.Moreover, for any given functionu: Ω→Rwe define uξ,y: Ωξ,y →Rbyuξ,y(t) =u(y+tξ). For the results collected in the following Theorem see [3], section 3.11.

Theorem 2.2. Let u∈ BV(Ω). Then uξ,y ∈ BV(Ωξ,y) for every ξ ∈ Sn−1 and for Hn−1-a.e.

y∈ξ. For such values ofy we haveu0ξ,y(t) =h∇u(y+tξ), ξifor a.e.t∈Ωξ,y andJuξ,y = (Ju)ξ,y, whereu0ξ,y denotes the absolutely continuous part of the measure derivative ofuξ,y. Moreover, for every open subsetA ofΩwe have

Z

ξ

|Dcuξ,y|(Aξ,y) dHn−1(y) =|hDcu, ξi|(A).

2.3. Γ-convergence. For the general theory see [9] and [18]. Let (X, d) be a metric space. Let (Fj) be a sequence of functionsX →R. We say that (Fj) Γ-converges, asj→+∞, toF:X→R, if for allu∈X we have:

a) For every sequence (uj) converging touit holds F(u)≤lim inf

j→+∞Fj(uj).

b) There exists a sequence (uj) converging to usuch that F(u)≥lim sup

j→+∞

Fj(uj).

ThelowerandupperΓ-limits of (Fj) inu∈X are defined as F0(u) = inf

lim inf

j→+∞Fj(uj) : uj→u , F00(u) = inf lim sup

j→+∞

Fj(uj) : uj→u

respectively. We extend this definition of convergence to families depending on a real parameter.

Given a family (Fε)ε>0 of functionsX →R, we say that it Γ-converges, asε→0, toF:X →R if for every positive infinitesimal sequence (εj) the sequence (Fεj) Γ-converges toF. If we define the lower and upper Γ-limits of (Fε) as

F0(u) = inf lim inf

ε→0 Fε(uε) : uε→u , F00(u) = inf lim sup

ε→0

Fε(uε) : uε→u

respectively, then (Fε) Γ-converges to F in uif and only ifF0(u) =F00(u) = F(u). It turns out that both F0 and F00 are lower semicontinuous on X. In the estimate of F0 we shall use the following immediate consequence of the definition:

F0(u) = inf lim inf

j→+∞Fεj(uj) : εj→0+, uj→u . It turns out that the infimum is attained.

An important consequence of the definition of Γ-convergence is the following result about the convergence of minimizers (see, e.g., [18], Cor. 7.20):

Theorem 2.3. LetFj:X→Rbe a sequence of functions whichΓ-converges to someF:X →R; assume thatinfv∈XFj(v)>−∞for everyj. Let(σj)be a positive infinitesimal sequence, and for every j letuj ∈X be aσj-minimizer ofFj, i.e.

Fj(uj)≤ inf

v∈XFj(v) +σj.

Assume thatuj→ufor someu∈X. Thenuis a minimum point ofF, and F(u) = lim

j→+∞Fj(uj).

(8)

Remark 2.4. The following property is a direct consequence of the definition of Γ-convergence:

if Fε

→ FΓ then Fε+G→ FΓ +G wheneverG:X →Ris continuous.

2.4. Supremum of measures. In order to prove the Γ-liminf inequality we recall the following useful tool, which can be found in [8].

Lemma 2.5. LetΩbe an open subset ofRnand denote byA(Ω)the family of its open subsets. Let λbe a positive Borel measure onΩ, andµ:A(Ω)→[0,+∞)a set function which is superadditive on open sets with disjoint compact closures, i.e. ifA, B⊂⊂ΩandA∩B=∅, then

µ(A∪B)≥µ(A) +µ(B).

Let (ψi)i∈I be a family of positive Borel functions. Suppose that µ(A)≥

Z

A

ψidλ for every A∈ A(Ω) andi∈I.

Then

µ(A)≥ Z

A

sup

i

ψidλ for everyA∈ A(Ω).

2.5. A density result. The right bound for the upper Γ-limit from above will be first obtained for a suitable dense subset of SBV(Ω). More precisely, let W(Ω) be the space of all functions w∈SBV(Ω) such that

(a) Hn−1(Sw\Sw) = 0;

(b) Sw is the intersection of Ω with the union of a finite member of (n−1)-dimensional simplexes;

(c) w∈Wk,∞(Ω\Sw) for everyk∈N.

Theorem 3.1 in [17] gives us the density property ofW(Ω) we need; here SBV2(Ω) ={u∈SBV(Ω) :|∇u| ∈L2(Ω),Hn−1(Su)<+∞}.

Theorem 2.6. Assume that ∂Ω is Lipschitz. Let u∈SBV2(Ω)∩L(Ω). Then there exists a sequence (wh) in W(Ω) such thatwh →u strongly in L1(Ω), ∇wh→ ∇ustrongly in L2(Ω,Rn), withlim suph→+∞kwhk≤ kuk and such that

lim sup

h→+∞

Z

Swh

ψ(w+h, wh, νwh) dHn−1≤ Z

Su

ψ(u+, u, νu) dHn−1

for every upper semicontinuous function ψ such that ψ(a, b, ν) = ψ(b, a,−ν) whenever a, b ∈ R andν ∈Sn−1.

2.6. A relaxation result. To conclude this section we prove a relaxation result which will be used in the sequel. Recall that givenX be a topological space andF:X →R∪ {±∞}, therelaxed functionalofF, denoted byF, is the largest lower semicontinuous functional which is smaller than F.

Theorem 2.7. Letφ: [0,+∞)→[0,+∞)be a convex, non-decreasing and lower semicontinuous function withφ(0) = 0 and with

t→+∞lim φ(t)

t =c∈(0,+∞).

Let θ: [0,+∞)×Sn−1 →[0,+∞)be a lower semicontinuous function such that θ(s, ν)≤c0s for any (s, ν)∈[0,+∞)×Sn−1, for somec0>0. For anyA∈ A(Ω)let

F(u, A) =





 Z

A

φ(|∇u|) dx+ Z

Su∩A

θ(|u+−u|, νu) dHn−1 ifu∈SBV2(Ω)∩L(Ω)

+∞ otherwise inL1(Ω).

Then the relaxed functional of F with respect to the strong L1-topology satisfies F(u)≤

Z

φ(|∇u|) dx+ Z

Su

θ(|u+−u|, νu) dHn−1+c|Dcu|(Ω)

(9)

for any u∈BV(Ω).

Proof. Combining a standard convolution argument with a well known relaxation result (see, for instance, Th. 5.47 in [3]) we can say that the relaxed functional of

G(u, A) =





 Z

A

φ(|∇u|) dx ifu∈C1(Ω) +∞ otherwise inL1(Ω) is given by

G(u, A) =





 Z

A

φ(|∇u|) dx+c|Dsu|(A) ifu∈BV(Ω)

+∞ otherwise inL1(Ω).

SinceC1(Ω)⊆SBV2(Ω)∩L(Ω) then we get F(u, A)≤ G(u, A). Hence for anyA∈ A(Ω) and for anyu∈BV(Ω)

F(u, A)≤ Z

A

φ(|∇u|) dx+c|Dsu|(A).

We can now conclude using the fact that for everyu∈BV(Ω) the set functionF(u,·) is the trace onA(Ω) of a regular Borel measureµ. This can be proven exactly along the same line of Prop. 3.3 in [6]. Hence

F(u) =µ(Ω) =µ(Ω\Su) +µ(Ω∩Su)

≤ Z

φ(|∇u|) dx+c|Dcu|(Ω) + Z

Su

θ(|u+−u|, νu) dHn−1

which is what we wanted to prove.

3. Statement of the main results

Let Ω⊂Rn be a bounded open set with Lipschitz boundary. Letφ: [0,+∞)→[0,+∞) be a convex and non-decreasing function withφ(0) = 0 and

(3.1) lim

t→+∞

φ(t)

t =c0∈(0,+∞).

For anyε >0 letfε: [0,+∞)→[0,+∞) be such that:

A1) fε is non-decreasing, continuous, withfε(0) = 0.

A2) It holds lim

(ε,t)→(0,0)

fε(t) εφ εt = 1.

A3) fε converges uniformly on the compact subsets of [0,+∞) to a concave functionf. Example 3.1. Given f andφas above, a possible choice for fε satisfying A1-A3 is given by

fε(t) =

 εφ εt

if0≤t≤tε

f(t−tε) +εφ tεε

ift > tε

wheretε→0, andtε/ε→+∞. The only non-trivial assumption to verify is A2. Sinceε/tφ(t/ε)→ c0 as(ε, t)→(0,0), witht≥tε, the check amounts to verify that

lim

(ε,t)→(0,0) t≥tε

f(t−tε) +εφ tεε

t =c0.

This follows immediately fromf(t−tε)/(t−tε)→c0 andε/tεφ(tε/ε)→c0 as(ε, t)→(0,0), and t≥tε.

(10)

LetC⊂Rnbe open, bounded, and connected with 0∈C. Letρ:C→(0,+∞) be a continuous and bounded convolution kernel with

Z

C

ρdx= 1.

For anyε >0 and for anyx∈Rn we will denote byCε(x) the setx+εC. For anyx∈εC let ρε(x) = 1

εnρx ε

.

We consider the family (Fε)ε>0 of functionalsL1(Ω)→[0,+∞] defined by

(3.2) Fε(u) =





 1 ε

Z

fε(ε|∇u| ∗ρε) dx if u∈W1,1(Ω)

+∞ otherwise inL1(Ω)

where, for anyx∈Ω,

(3.3) |∇u| ∗ρε(x) =

Z

Cε(x)∩Ω

|∇u(y)|ρε(y−x) dy is a regularization by convolution of|∇u|by means of the kernel ρε. Remark 3.2. Notice that with the choiceC=B1 andρ= ω1

n

χB

1 we get

|∇u| ∗ρε(x) = Z

Bε(x)∩Ω

|∇u|dy

and thus the family(Fε)ε>0 reduces to the case already investigated in[20],[21] and[22].

In order to prove the Γ-convergence of Fε it is convenient to introduce a localized version of Fε: more precisely, for eachA∈ A(Ω) we set

(3.4) Fε(u, A) =





 1 ε

Z

A

fε(ε|∇u| ∗ρε) dx if u∈W1,1(Ω)

+∞ otherwise inL1(Ω).

Clearly,Fε ·,Ω

coincides with the functionalFεdefined in (3.2). The lower and upper Γ-limits of Fε(·, A)

will be denoted byF0(·, A) andF00(·, A), respectively.

3.1. The anisotropy. In this paragraph we define the surface density θ: [0,+∞)×Sn−1→[0,+∞) which will appear in the expression of the Γ-limit ofFε.

Givenν ∈Sn−1 anda, b∈Rlet us denote byua,bν the functionRn→Rgiven by ua,bν (x) =

a ifhx, νi<0 b ifhx, νi ≥0.

For anyx∈Rn and anyν ∈Sn−1 letPν(x) be the orthogonal projection ofxonto the subspace ν ={x∈Rn :hx, νi= 0}. We define the cylinder

Qν ={x∈Rn:|hx, νi| ≤1, Pν(x)∈B1∩ν}.

Given Ω0 ⊂Rn with Qν ⊂⊂Ω0 denote by Wνa,b the space of all sequences (uj) inWloc1,1(Ω0) such that uj →ua,bν in L1(Ω0), and such that there exist two positive infinitesimal sequences (aj),(bj) withuj(x) =aifhx, νi<−aj anduj=b ifhx, νi> bj. Let

(3.5) θ(s, ν) = 1 ωn−1inf

lim inf

j→+∞

1 εj

Z

Qν

f(εj|∇uj| ∗ρεj) dx: (uj)∈Wν0,s, εj →0+

. Notice thatθ(s, ν) does not depend on the choice of Ω0. Let us collect some easy properties ofθ which immediately descend from the definition.

(11)

Lemma 3.3. The following properties hold:

(3.6) θ is continuous.

(3.7) θ(s, ν) =θ(s,−ν), ∀s≥0, ∀ν ∈Sn−1.

(3.8)

inf

lim inf

j→+∞

1 εj

Z

Qν

f(εj|∇uj| ∗ρεj) dx: (uj)∈Wν0,s, εj →0+

= inf

lim inf

j→+∞

1 εj

Z

Qν

f(εj|∇uj| ∗ρεj) dx: (uj)∈Wνa,b, εj →0+

whenever|a−b|=s.

Moreover, for any x0∈Rn,ν ∈Sn−1 ands≥0 we have (3.9) θ(s, ν) = 1

ωn−1inf

lim inf

j→+∞

1 εj

Z

x0+Qν

f(εj|∇uj| ∗ρεj) dx: (uj(· −x0))∈Wν0,s, εj→0+

.

3.2. Main results. We are now in position to state the main result of the paper.

Theorem 3.4. Let Fεbe as in (3.2), withfε satisfying conditions A1-A3. ThenFεΓ-converges, with respect to the strong L1-topology, asε→0, toF: L1(Ω)→[0,+∞]given by

F(u) =





 Z

φ(|∇u|) dx+ Z

Su

θ(|u+−u|, νu) dHn−1+c0|Dcu|(Ω) if u∈GBV(Ω)

+∞ otherwise in L1(Ω).

Remark 3.5. Notice that for any u∈GBV(Ω) the expression θ(|u+−u|, νu)turns out to be well definedHn−1-a.e.x∈Su, since (3.7)holds.

The proof of Theorem 3.4 will descend combining Proposition 5.10 (the Γ-liminf inequality) with Proposition 6.3 (the Γ-limsup inequality).

As a typical consequence of a Γ-convergence result, we are able to prove a result of convergence of minima by means of the following compactness result for equibounded (in energy) sequences, which will be proved in§4.

Theorem 3.6. Let (εj)be a positive infinitesimal sequence, and let (uj)be a sequence inL1(Ω) such that ||uj|| ≤M, and such thatFεj(uj)≤M for some positive constantM independent of j. Then the sequence(uj)converges, up to a subsequence, in L1(Ω) to a function u∈BV(Ω).

Theorem 3.7. Let (εj) be a positive infinitesimal sequence and let g ∈L(Ω). For everyu ∈ L1(Ω) andj∈Nlet

Ij(u) =Fεj(u) + Z

|u−g|dx, I(u) =F(u) + Z

|u−g|dx . For everyj letuj ∈L1(Ω) be such that

Ij(uj)≤ inf

L1(Ω)

Ijj.

Then the sequence(uj)converges, up to a subsequence, to a minimizer of I inL1(Ω).

Proof. Since g ∈ L(Ω) and since Fεj decreases by truncation, we can assume that (uj) is equibounded in L(Ω); for instance ||uj|| ≤ ||g||. Applying Theorem 3.6 there exists u ∈ BV(Ω) such that (up to a subsequence)uj→uinL1(Ω). By Theorem 2.3, since (Ij) Γ-converges toI (see Th. 3.4 and Remark 2.4), uis a minimum point of I onL1(Ω).

(12)

4. Compactness

In this section we prove Theorem 3.6. Let us first recall a useful technical Lemma which can be found in [10], Prop. 4.1. Actually such a Proposition has been proved for|∇u|2, but, up to simple modifications, the same proof works for|∇u|.

For everyA∈ A(Ω) andσ >0 we set

Aσ={x∈A: d(x, ∂A)> σ}.

Lemma 4.1. Let g: [0,+∞)→[0,+∞)be a non-decreasing continuous function such that

t→0lim g(t)

t =c

for some c >0. LetA∈ A(Ω) withA⊂⊂Ω, and let u∈W1,1(Ω)∩L(Ω). For anyδ >0 and for any ε >0 sufficiently small, there exists a function v∈SBV(A)∩L(A) such that

(1−δ) Z

A

|∇v|dx≤1 ε

Z

A

g

ε Z

Bε(x)

|∇u|dy

dx,

Hn−1(Sv∩A)≤c0 ε

Z

A

g

ε Z

Bε(x)

|∇u|dy

dx, kvkL(A)≤ kukL(A)

kv−ukL1(A)≤c0kukL(A)

Z

A

g

ε Z

Bε(x)

|∇u|dy

dx, wherec0 is a constant depending only onn, δ andg.

Proof of Theorem 3.6. LetA∈ A(Ω) withA⊂⊂Ω and∂Asmooth. Letr >0 such thatBr⊂C, and letm= infBrρ >0. Then for anyx∈Awe haveBj(x)⊂Cεj(x) and thus forj sufficiently large,

|∇uj| ∗ρεj(x) = Z

Cεj(x)

|∇uj(y)|ρεj(y−x) dy≥ m εnj

Z

Brεj(x)

|∇uj(y)|dy

=mrnωn

Z

Brεj(x)

|∇uj(y)|dy

for any x∈ A. Fixδ > 0. By A2 there exist tδ >0 and jδ such thatfεj(t)≥(1−δ)εjφ(t/εj) for anyt ∈[0, tδ] and j > jδ. Letα, β ∈R, withα > 0 and β <0, be such that φ(t)≥αt+β everywhere. Then, sincefεj is non-decreasing, we havefεj(t)≥gεδ

j(t) for anyt≥0, being gδε

j(t) =

(1−δ)αt+εjβ ift∈[0, tδ] (1−δ)αtδjβ ift > tδ. Therefore, lettinghδ(t) =gεδj(t)−εjβ, we have

(4.1)

Fεj(uj, A)≥ 1 εj

Z

A

hδ(|∇uj| ∗ρεj) dx+β|A|

≥ 1 εj

Z

A

hδ

mrnωnεj

Z

Brεj(x)

|∇uj|dy

dx+β|A|.

Letηj =rεj andgδ,m,r(t) =1rgδ(mrn−1ωnt).Notice that, by construction,

t→0lim

gδ,m,r(t) t exists and is finite. Then inequality (4.1) becomes

Fεj(uj, A)−β|A| ≥ 1 ηj

Z

gδ,r,m

ηj

Z

Bηj(x)

|∇uj|dy

dx.

(13)

Applying Lemma 4.1 we find a sequence (vj) inSBV(A) and a constantCindependent ofAsuch thatkvjkBV(A)≤C andkvjkL(A)≤C.Moreover,

(4.2) kvj−ujkL1(A)→0.

Hence, by Theorem 2.1, the sequence (vj) converges, up to a subsequence not relabeled, to some u ∈ BV(A), with kukBV(A) ≤ C. By (4.2) also uj converges to u in L1(A). The arbitrariness ofA and a diagonal argument allow to find a subsequence (ujk) which converges in L1loc(Ω) to a functionu∈BVloc(Ω), and the uniform bound of kujkL(Ω)implies the convergence is strong in

L1(Ω).

5. TheΓ-liminf inequality In this section we will prove that for anyu∈L1(Ω) the inequality

F(u)≤lim inf

j→+∞Fεj(uj)

holds for anyuj→uinL1(Ω). First we will investigate two particular situations.

5.1. A preliminary estimate from below in terms of the volume and Cantor parts. In this paragraph we will take into account a simpler family of functionals. Let α, β > 0 and let g: [0,+∞)→[0,+∞) given byg(t) =αt∧β. LetGε:L1(Ω)× A(Ω)→[0,+∞] be defined by

Gε(u, A) =





 1 ε

Z

A

g(ε|∇u| ∗ρε) dx if u∈W1,1(Ω)

+∞ otherwise inL1(Ω).

We wish to estimate from below the lower Γ-limitG0(·, A) in terms of the volume and the Cantor parts ofDu. To this sake, we apply a slicing procedure, so that at first we will establish a suitable one-dimensional inequality. The idea of the proof is the same as in [25], where the superlinear growth case is treated.

Letm∈Nodd, letAbe an open interval inR, and let (εj) be a positive infinitesimal sequence.

LetAj={x∈εjZ:x∈A}.For anyj∈Nand for anyx∈Aj we define the interval Ij(x) =h

x−mεj

2 , x+mεj 2

i .

Lemma 5.1. Let α0, β0 > 0 and let hj: [0,+∞) → [0,+∞) given by hj(t) = α0t∧ βε0

j. Let u∈BV(A)and letuj→uinL1(A) withuj∈W1,1(A)for anyj∈N. Then

(5.1) lim inf

j→+∞εj X

x∈Aj

hj Z

Ij(x)

|u0j|dy

≥α0 Z

A

|u0|dy+α0|Dcu|(A).

Proof. For anyj∈Nandi= 0, . . . , m−1 letAij = (iεj+mεjZ)∩A.ObviouslyAj is the disjoint union ofAij fori∈ {0, . . . , m−1}. Then

X

x∈Aj

hj Z

Ij(x)

|u0j|dy

≥ 1 m

m−1

X

i=0

X

x∈Aij

mhj Z

Ij(x)

|u0j|dy

.

Now let

Aij = (

x∈Aij: Z

Ij(x)

|u0j|dx≤ β0 α0εj

)

and letvj∈SBV(A) given by vj(x) =

uj(x) ifx∈S

y∈AijIj(y) 0 otherwise inA.

(14)

Hence X

x∈Aij

jhj Z

Ij(x)

|u0j|dy

≥ X

x∈Aij

jhj Z

Ij(x)

|u0j|dy

0 X

x∈Aij

Z

Ij(x)

|u0j|dy

0 Z

A

|v0j|dy.

Observe that since we can suppose, without loss of generality, that εj X

x∈Aj

hj Z

Ij(x)

|u0j|dy

≤M

for someM ≥0, we deduce that M ≥εj

X

x∈Aj\Sm−1 i=0 Aij

hj

Z

Ij(x)

|u0j|dy

j

β0 εj

]

Aj\

m−1

[

i=0

Aij

from which necessarily we have εj]

Aj\

m−1

[

i=0

Aij

→0, asj →+∞.

This implies that ||uj−vj||L1(A) →0 asj →+∞. Therefore, vj →uin L1(A). Finally, by the superadditivity of the lim inf and by the lower semicontinuity of the total variation, we get

lim inf

j→+∞εj X

x∈Aj

hj Z

Ij(x)

|u0j|dy

≥ 1 m

m−1

X

i=0

lim inf

j→+∞

X

x∈Aij

jhj Z

Ij(x)

|u0j|dy

≥α0lim inf

j→+∞

Z

A

|vj0|dy≥α0|Du|(A)

≥α0 Z

A

|u0|dy+α0|Dcu|(A)

which ends the proof.

Now, by applying the slicing Theorem 2.2, we will reduce the n-dimensional inequality to the one-dimensional inequality 5.1. Fix ξ∈Sn−1 and δ∈(0,1); consider an orthonormal basis {ei} withen=ξ. Let

Qξδ =

x∈Rn:|hx,eii| ≤ δ

2, i= 1, . . . , n

, Qξδ(x) =x+Qξδ

and the lattice Zδξ = {x ∈ Rn : hx,eii ∈ δZ, i = 1, . . . , n}. In what follows we will denote by gj(t) =ε1

jg(εjt); in particular it holds gj(t) =αt∧ εβ

j and Gεj(u, A) =

Z

A

gj(|∇u| ∗ρεj) dx, u∈W1,1(Ω).

Finally fix A∈ A(Ω) and letAξδ ={x∈ Zδξ : Qξδ(x) ⊂A}. The following Lemma is a standard easy application of the mean value Theorem (see also Lemma 4.2 in [10]).

Lemma 5.2. Let u∈W1,1(Ω). Then there existsτ ∈Qξδ such that Gεj(u, A)≥ X

x∈Aξδ

δngj(|∇u| ∗ρεj(x+τ)).

Proof. We have

Gεj(u, A)≥ X

x∈Aξδ

Z

Qξδ(x)

gj(|∇u| ∗ρεj(y)) dy= Z

Qξδ

X

x∈Aξδ

gj(|∇u| ∗ρεj(y+x)) dy.

(15)

Applying the mean value Theorem we get Z

Qξδ

X

x∈Aξδ

gj(|∇u| ∗ρεj(y+x)) dy= X

x∈Aξδ

gj(|∇u| ∗ρεj(τ+x))

for someτ ∈Qξδ, which concludes the proof.

We are in position to apply the slicing procedure.

Proposition 5.3. Let u∈BV(Ω) andA∈ A(Ω). Then G0(u, A)≥α

Z

A

|∇u|dx and G0(u, A)≥α|Dcu|(A).

Proof. Fixξ∈Sn−1. For anyη >0 letPηξbe the union of the squaresQξη(yi)⊂Cwithyi∈Zηξ for i= 1, . . . , m, for somem∈Ndepending onη andξ. Let ρη be a non-negative constant function on the squaresQξη(yi) with 0< ρη≤ρand such that

cη= Z

C

ρηdx→1, as η→0.

Letciη(yi); then we can rewrite cη as cη =Pm

i=1ciηn.LetPηεξ

j be the union of the squares Qξηεj(yi)⊆Cεj, withyi ∈Zηεξ j, fori= 1, . . . , m. LetAξj =Aξηεj; applying Lemma 5.2, since we can suppose, without loss of generality, thatuj∈W1,1(Ω), there existsτj ∈Qξηεj such that

Gεj(uj, A)≥ X

x∈Aξj

(ηεj)ngj(|∇uj| ∗ρεj(x+τj)).

LetB ⊂⊂A, and, for anyj sufficiently large, letvj(y) =uj(y+τj). Then we getvj ∈W1,1(B) andvj →uin L1(B). Thus

Gεj(uj, A)≥ X

x∈Bjξ

(ηεj)ng(|∇vj| ∗ρεj(x))

being Bjξ ={x∈Zηεξ j : Qξηεj ⊆B}. Now, for eachx∈Bjξ, we estimate the term|∇vj| ∗ρεj(x);

we have, forj large enough,

|∇vj| ∗ρεj(x) = Z

Cεj

|∇vj(y+x)|ρεj(y) dy≥ 1 εnj

Z

Pηεjξ

|∇vj(y+x)|ρη

y εj

dy

≥ 1 εnj

m

X

i=1

ci

Z

Qξηεj(yi)

|∇vj(y+x)|dy=

m

X

i=1

ciηn cη

Z

Qξηεj(yi)

cη|∇vj(y+x)|dy.

SincePm i=1

ciηn

cη = 1 and sincegj is concave we get, for everyx∈Bjξ, gj(|∇vj| ∗ρεj(x))≥

m

X

i=1

ciηn cη gj

cη

Z

Qξηεj(yi)

|∇vj(y+x)|dy

.

Thus, reordering the terms, we deduce that Gεj(uj, A)≥ X

x∈Djξ

(ηεj)ngj

cη

Z

Qξηεj(x)

|∇vj|dz

for any D ⊂⊂ B and j sufficiently large, being, as usual, Dξj = {x ∈ Zηεξ j : Qξηεj ⊆ D}. For convenience we can suppose∇vj = 0 on

Rn\ [

Qξηεj⊆D

Qξηεj.

(16)

Let hξi be the one-dimensional space generated by ξ. Let us denote by Zηεξkj and by Zηεξj the orthogonal projections ofZηεξ j respectively on hξiand ξ. Then

Gεj(uj, A)≥ X

x∈Zηεjξ

(ηεj)ngj

cη Z

Qξηεj(x)

|∇vj|dz

≥ X

x∈Zξηεj

X

xk∈Zξηεjk

(ηεj)ngj

cη

Z

Qξηεj(x+xk)

|∇vj|dz

wherex=xk+x turns out to be the unique decomposition of anyx∈Zηεξ j withxk∈Zηεξkj and x ∈Zηεξj. Moreover, denoting by Qξηεkj and by Qξηεj the projections of Qξηεj respectively on hξi and onξ, applying Jensen’s inequality we deduce that

Gεj(uj, A)≥ X

x∈Zηεjξ

X

xk∈Zηεjξk

(ηεj)ngj

cη

Z

Qξηεj(x)

Z

Qξηεjk (xk)

|h∇vj(z+zk), ξi|dzkdz

≥ X

x∈Zηεjξ

X

xk∈Zηεjξk

(ηεj)n Z

Qξηεj(x)

gj

cη

Z

Qξηεjk (xk)

|h∇vj(z+zk), ξi|dzk

dz

≥ X

x∈Zηεjξ

Z

Qξηεj(x)

X

xk∈Zηεjξk

ηεjgj

cη

Z

Qξηεjk (xk)

|h∇vj(z+zk), ξi|dzk

dz

≥ Z

ξ

X

xk∈Zηεjξk

ηεjgj

cη

Z

Qξηεjk (xk)

|h∇vj(z+zk), ξi|dzk

dz.

For anyσ >0 small letDσ ={x∈D:d(x, ∂D)> σ} andDxσ ={x∈Dσ :x=x+xkξ, xk ∈ R}, for x ∈ ξ. For j sufficiently large, vj(x+·) ∈ W1,1(Dxσ). Furthermore, vj → u in L1(Dxσ) for a.e.x∈ξ. Lethj(t) =gj(cηt); then, by the very definition of g, it is easy to see thathj(t) =αcηt∧εβ

j.We are in position to apply Lemma 5.1 with choice α0 =αcη andβ0=β.

Thus

lim inf

j→+∞

X

xk∈Zξηεjk

ηεjgj

cη Z

Qξηεjk (xk)

|h∇vj(z+zk), ξi|dzk

= lim inf

j→+∞

X

xk∈Zξηεjk

ηεjhj Z

Qξηεjk (xk)

|h∇vj(z+zk), ξi|dzk

≥αcη

Z

Dσz

|h∇u(z+zk), ξi|dzk+αcη|hDcu(z+·), ξi|(Dzσ).

Taking into account Theorem 2.2 and Fatou’s Lemma we conclude that lim inf

j→+∞Gεj(uj, A)≥cηα Z

Dσ

|h∇u(z), ξi|dz+cηα|hDcu, ξi(Dσ).

Sincecη →1 asη→0, letσ→0 andD%A. Then (5.2) G0(u, A)≥α

Z

A

|h∇u(z), ξi|dz and G0(u, A)≥α|hDcu, ξi|(A)

for any ξ ∈ Sn−1. From the first inequality, using the superadditivity of G0 and Lemma 2.5 we easily deduce that

G0(u, A)≥α Z

A

|∇u|dz.

Referenzen

ÄHNLICHE DOKUMENTE

Let us begin by recalling that the conformal block decomposition of the s- and t- channel Polyakov-Regge blocks with conformal dimension ∆ and spin J have the same structure as

For each optimal center o ∈ O, let s o denote its closest heuristic center in S. Note that each optimal center is captured by exactly one heuristic center, but each heuristic center

By adopting the GW 0 CBM energy into the F band emission energy in Table V, we find that the lattice relaxation in presence of the electron-phonon interac- tion is best accounted by

Payment uncertainty is one of the areas in which MIRBAP scores better than conventional result-based payments – since payments are based on ex-ante results of modelling, payment is

Although we conjecture that the energy space of E τ is characterized by sets with local graph representations satisfying Lipschitz continuity and a suitable fractional

Examples. Use the introduction axioms. It suffices to prove the premises of Even − for this X. Consider the algebra Y of binary trees. We want to represent the set of total ideals by

Therefore, in the following local hybrid functionals based on advanced exchange and correlation functionals (GGA-based local hybrids), along with calibrated exchange- energy

INTRODUCTION 2 error estimators, we have to solve the dual problem, but the dual problem is still linear even if the problem is not linear (see [12]), and, therefore, for