Similar documents
Γ -convergence of the Allen Cahn energy with an oscillating forcing term

SOME MEASURABILITY AND CONTINUITY PROPERTIES OF ARBITRARY REAL FUNCTIONS

Coupled second order singular perturbations for phase transitions

CHAPTER 6. Differentiation

Rearrangements and polar factorisation of countably degenerate functions G.R. Burton, School of Mathematical Sciences, University of Bath, Claverton D

Global minimization. Chapter Upper and lower bounds

Stanford Mathematics Department Math 205A Lecture Supplement #4 Borel Regular & Radon Measures

ON TRIVIAL GRADIENT YOUNG MEASURES BAISHENG YAN Abstract. We give a condition on a closed set K of real nm matrices which ensures that any W 1 p -grad

Pointwise convergence rate for nonlinear conservation. Eitan Tadmor and Tao Tang


NECESSARY CONDITIONS FOR WEIGHTED POINTWISE HARDY INEQUALITIES

Projektbereich A. of Behavioral Heterogeneity. Kurt Hildenbrand ) July 1998

Linear Algebra (part 1) : Vector Spaces (by Evan Dummit, 2017, v. 1.07) 1.1 The Formal Denition of a Vector Space

Contents: 1. Minimization. 2. The theorem of Lions-Stampacchia for variational inequalities. 3. Γ -Convergence. 4. Duality mapping.

NONTRIVIAL SOLUTIONS TO INTEGRAL AND DIFFERENTIAL EQUATIONS

GIOVANNI COMI AND MONICA TORRES

2 BAISHENG YAN When L =,it is easily seen that the set K = coincides with the set of conformal matrices, that is, K = = fr j 0 R 2 SO(n)g: Weakly L-qu

g 2 (x) (1/3)M 1 = (1/3)(2/3)M.

2. Function spaces and approximation

c Birkhauser Verlag, Basel 1997 GAFA Geometric And Functional Analysis

Tools from Lebesgue integration

The Erwin Schrodinger International Boltzmanngasse 9. Institute for Mathematical Physics A-1090 Wien, Austria

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

Math The Laplacian. 1 Green s Identities, Fundamental Solution

ON MAPS WHOSE DISTRIBUTIONAL JACOBIAN IS A MEASURE

Some Background Material

Green s Functions and Distributions

JUHA KINNUNEN. Harmonic Analysis

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

at time t, in dimension d. The index i varies in a countable set I. We call configuration the family, denoted generically by Φ: U (x i (t) x j (t))

The Dirichlet s P rinciple. In this lecture we discuss an alternative formulation of the Dirichlet problem for the Laplace equation:

Integral Jensen inequality

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

On a weighted total variation minimization problem

Real Analysis Notes. Thomas Goller

i=1 α i. Given an m-times continuously

1. Introduction Boundary estimates for the second derivatives of the solution to the Dirichlet problem for the Monge-Ampere equation

2 Garrett: `A Good Spectral Theorem' 1. von Neumann algebras, density theorem The commutant of a subring S of a ring R is S 0 = fr 2 R : rs = sr; 8s 2


Introduction to Real Analysis Alternative Chapter 1

Minimum and maximum values *

pacific journal of mathematics Vol. 181, No. 2, 1997 MOMENT MAPS ON SYMPLECTIC CONES Suzana Falc~ao B. de Moraes and Carlos Tomei We analyze some conv

Convex Analysis and Optimization Chapter 2 Solutions

CHAPTER 9. Embedding theorems

B. Appendix B. Topological vector spaces

2 Sequences, Continuity, and Limits

1 Directional Derivatives and Differentiability

Part III. 10 Topological Space Basics. Topological Spaces

Riemann integral and volume are generalized to unbounded functions and sets. is an admissible set, and its volume is a Riemann integral, 1l E,

Topological properties

On non negative solutions of some quasilinear elliptic inequalities

Behaviour of Lipschitz functions on negligible sets. Non-differentiability in R. Outline

Laplace s Equation. Chapter Mean Value Formulas

Set, functions and Euclidean space. Seungjin Han

Real Analysis Problems

Calculus of Variations. Final Examination

1 Solutions to selected problems

THE LUBRICATION APPROXIMATION FOR THIN VISCOUS FILMS: REGULARITY AND LONG TIME BEHAVIOR OF WEAK SOLUTIONS A.L. BERTOZZI AND M. PUGH.

NOTES ON THE REGULARITY OF QUASICONFORMAL HOMEOMORPHISMS

(convex combination!). Use convexity of f and multiply by the common denominator to get. Interchanging the role of x and y, we obtain that f is ( 2M ε

Congurations of periodic orbits for equations with delayed positive feedback

MAT 257, Handout 13: December 5-7, 2011.

Measure and Integration: Solutions of CW2

r( = f 2 L 2 (1.2) iku)! 0 as r!1: (1.3) It was shown in book [7] that if f is assumed to be the restriction of a function in C

Problem set 1, Real Analysis I, Spring, 2015.

PARAMETER IDENTIFICATION IN THE FREQUENCY DOMAIN. H.T. Banks and Yun Wang. Center for Research in Scientic Computation

QUADRATIC RATE OF CONVERGENCE FOR CURVATURE DEPENDENT SMOOTH INTERFACES: A SIMPLE PROOF 1

Unstable K=K T=T. Unstable K=K T=T

Lecture 8: Basic convex analysis

2 JOSE BURILLO It was proved by Thurston [2, Ch.8], using geometric methods, and by Gersten [3], using combinatorial methods, that the integral 3-dime

On Coarse Geometry and Coarse Embeddability

1 Topology Definition of a topology Basis (Base) of a topology The subspace topology & the product topology on X Y 3

AN EXAMPLE OF FUNCTIONAL WHICH IS WEAKLY LOWER SEMICONTINUOUS ON W 1,p FOR EVERY p > 2 BUT NOT ON H0

M311 Functions of Several Variables. CHAPTER 1. Continuity CHAPTER 2. The Bolzano Weierstrass Theorem and Compact Sets CHAPTER 3.

Chapter 7. Extremal Problems. 7.1 Extrema and Local Extrema

C 1 convex extensions of 1-jets from arbitrary subsets of R n, and applications

INTRODUCTION TO NETS. limits to coincide, since it can be deduced: i.e. x

ESTIMATES FOR THE MONGE-AMPERE EQUATION

Sobolev Spaces. Chapter Hölder spaces

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

CHAPTER I THE RIESZ REPRESENTATION THEOREM

Abstract. Jacobi curves are far going generalizations of the spaces of \Jacobi

FLUX IDENTIFICATION IN CONSERVATION LAWS 2 It is quite natural to formulate this problem more or less like an optimal control problem: for any functio

A Finite Element Method for an Ill-Posed Problem. Martin-Luther-Universitat, Fachbereich Mathematik/Informatik,Postfach 8, D Halle, Abstract

Stochastic dominance with imprecise information

Techinical Proofs for Nonlinear Learning using Local Coordinate Coding

quantitative information on the error caused by using the solution of the linear problem to describe the response of the elastic material on a corner

SHARP BOUNDARY TRACE INEQUALITIES. 1. Introduction

Variational Formulations

CONSTRAINED PERCOLATION ON Z 2

Hamburger Beiträge zur Angewandten Mathematik

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

Contractive metrics for scalar conservation laws

Second Order Elliptic PDE

Convex Analysis and Economic Theory Winter 2018

MATH 51H Section 4. October 16, Recall what it means for a function between metric spaces to be continuous:

1. Bounded linear maps. A linear map T : E F of real Banach


Transcription:

Max-Planck-Institut fur Mathematik in den Naturwissenschaften Leipzig A nonlocal anisotropic model for phase transitions Part II: Asymptotic behaviour of rescaled energies by Giovanni Alberti and Giovanni Bellettini Preprint-Nr.: 9 1997

A Nonlocal Anisotropic Model for Phase Transitions Part II: Asymptotic Behaviour of Rescaled Energies Giovanni Alberti* Giovanni Bellettini** Abstract: we study the asymptotic behaviour as! 0, of the nonlocal models for phase transition described by the scaled free energy F (u) := 1 4 J (x 0? x)? u(x 0 )? u(x) 2 dx 0 dx + 1 W? u(x) dx ; where u is a scalar density function, W is a double-well potential which vanishes at 1, J is a non-negative interaction potential and J (h) :=?N J(h=). We prove that the functionals F converge in a variational sense to the anisotropic surface energy F (u) := Su ( u ) ; where u is allowed to take the values 1 only, u is the normal to the interface Su between the phases fu = +1g and fu =?1g, and is the surface tension. This paper concludes the analisys started in [AB]. Keywords: phase transitions, singular perturbations,?-convergence, nonlocal integral functionals. 1991 AMS Subject Classication: 49J45, 49N45, 82B24 1. Introduction and statement of the result In this paper we study the asymptotic behaviour as! 0 of the functionals F dened by F (u; ) := 1 4 J (x 0? x)? u(x 0 )? u(x) 2 dx 0 dx + 1 W? u(x) dx ; (1:1) * Max-Planck-Institut fur Mathematik, Inselstr. 22-26, D-04103 Leipzig (Germany) on leave from Dipartimento di Matematica Applicata, Universita di Pisa (Italy) e-mail: alberti@mis.mpg.de ** Dipartimento di Matematica Applicata \U. Dini, via Bonanno 25/B, I-56126 Pisa (Italy) e-mail: belletti@dm.unipi.it 1

and of their minimizers subject to suitable constraints. Here u is a real function on the open set R N, is a positive scaling parameter, J : R N! [0; +1] is an even function and J (y) :=?N J(y=), W : R! [0; +1) is a continuous function which vanishes at 1 only (see paragraph 1.2 for precise assumptions). The function u can be interpreted as a macroscopic density of a scalar intrinsic quantity which describes the congurations of a given system; the second integral in (1:1) forces a minimizing conguration u to take values close to the \pure states +1 and?1 (phase separation), while the rst integral represents an interaction energy which penalizes the spatial inhomogeneity of the conguration (surface tension). A relevant example is given in equilibrium Statistical Mechanics by the continuum limit of Ising spin systems on lattices; in that setting u represents a macroscopic magnetization density and J is a ferromagnetic Kac potential (cf. [ABCP] and references therein). R If the parameter is small, when we minimize F subject to the volume constraint u = c with jcj < vol(), then the second term in (1:1) prevails; roughly speaking a minimizer u takes values close to?1 or +1, and the transition between the two phases occurs in a thin layer with thickness of order. It is therefore natural to consider the asymptotic behaviour of this model as tends to 0; accordingly we expect that the minimizers u converge (possibily passing to a subsequence) to a limit function u which takes values 1 only. More precisely our main result states the following (see Theorem 1.4 and remarks below): as! 0 the functionals F converge, in the sense of?-convergence in L 1 (), to a limit energy F which is nite only when u = 1 a.e., and in that case is given by F (u) := Su ( u ) dh N?1 ; (1:2) where Su is the interface between the phases fu = +1g and fu =?1g, u if the normal eld to Su and is a suitable strictly positive even function on R N ; H N?1 denotes the (N? 1)- dimensional Hausdo measure. The denition of?-convergence immediately implies that the minimizers u of F converge in L 1 () to minimizers of F. This means that when! 0 the model associated with the energy F converge to the classical van der Waals model for phase separation associated with the (anisotropic) surface tension. A rst result of this type was proved in the isotropic case (that is, when J is radially symmetric) in [ABCP], although with a dierent method and for a particular choice of W only. In the isotropic case F (u) reduces to the measure of the interface Su multiplied by a positive factor, which is obtained by taking a suitable (unscaled) one-dimensional functional F (v; R) of type (1:1) and computing its inmum over all v : R! [?1; 1] which tend to +1 at +1 and to?1 at?1. We show that in the general anisotropic case the value of the function at some unit vector e is obtained by solving a certain minimum problem on an unbounded N-dimensional stripe; this is called the optimal prole problem associated with the direction e (see paragraph 1.3). In [AB] it was proved that for every e such a problem admits at least one solution u : R N! [?1; 1] which is constant with respect to all directions orthogonal to e. We emphasize that the proof of the?-convergence theorem does not depend on this existence result (cf. Remark 1.10). Indeed our result can also be extended to the multi-phase case, that is, when u takes values in R m and W is a function on R m which vanishes at nitely many anely independent points (see paragraph 1.12), while so far there are no existence results for the corresponding optimal prole problem (cf. [AB], section 4b). Finally we underline that the assumption J 0, which in the statistical model is addressed as the ferromagnetic assumption, is crucial in all our proofs, while the other assumptions on J and W are close to be optimal (see paragraphs 1.2 and 1.13). 2

For a xed positive our model closely recall the Cahn-Hilliard model for phase separation (see [CH]), which is described by the energy functional I (u) := 2 jruj 2 + 1 W (u) : (1:3) Indeed F can be obtained from I by replacing the term ru(x) in the rst integral in (1:3) u(x + h)? u(x) with respect to the measure with the average of the nite dierences 1 distribution J(h) dh. The?-convergence of the functionals I to a limit (isotropic) energy of the form (1:2) was proved by L. Modica and S. Mortola (see [MM], [Mo]) and then extended to R more general anisotropic functionals in [Bou], [OS]. It is important to recall that the term jruj 2 in (1:3) was derived in [CH] as a rst order approximation of a more general and complicated quadratic form. Indeed our result shows that the asymptotic behaviour of these functionals is largely independent of the choice of this quadratic form. A rst result in this direction was already obtained in [ABS]; the one-dimensional functionals considered there, were dened as in (1:1) with J(h) = 1=h 2 (see also paragraph 1.13 below). We recall here that the evolution model associated with the energy F is described, after a suitable time scaling, by the nonlocal parabolic equation u t =?2? J u? u? f(u) ; (1:4) where f is the derivative of W and we assumed kjk 1 = 1; the analog for the energy I is the scaled Allen-Cahn equation u t = 4u??2 f(u). The asymptotic behaviour of the solutions of (1:4) has been widely studied in the isotropic case (see for instance [DOPT1-3], [KS1-2]) and leads to a motion by mean curvature in the sense of viscosity solutions; the generalization to the anisotropic case has been given in [KS3]. Analogous results have been proved for the solutions of the scaled Allen-Cahn equation (see for instance [BK], [DS], [ESS], [Ilm]). An interesting mathematical feature of the functional F is that they are not local; this means that given disjoint sets A and A 0, the energy F (u; A [ A 0 ) stored in A [ A 0 is strictly larger than the sum of F (u; A) and F (u; A 0 ). More precisely we have F (u; A [ A 0 ) = F (u; A) + F (u; A 0 ) + 2 (u; A; A 0 ) ; (1:5) where the locality defect (u; A; A 0 ) is dened as (u; A; A 0 ) := 1 4 AA 0 J (x 0? x)? u(x 0 )? u(x) 2 dx 0 dx (1:6) for every A; A 0 R N and every u : A [ A 0! R. The meaning of (u; A; A 0 ) is quite clear, since it represents the scaled interaction energy between A and A 0. In this paper we always assume a proper decay of J at innity, namely (1:8), in order to guarantee that (u; A; A 0 ) vanishes as! 0 whenever the sets A and A 0 are distant. Before passing to precise statements, we need to x some general notation. In the following is a bounded open subset of R N, and it is called regular when has Lipschitz boundary (for N = 1, when it is a nite union of distant open intervals). Unless dierently stated all sets and functions are assumed to be Borel measurable. Every set in R N is usually endowed with the Lebesgue measure L N, and we simply write R B f(x) dx for the integrals over B and jbj for L N(B), while we never omit an explicit mention 3

of the measure when it diers from L N. As usual H N?1 denotes the (N? 1)-dimensional Hausdor measure. 1.1. BV functions and sets of nite perimeter For every open set in R N, BV () denotes the space of all functions u :! R with bounded variation, that is, the functions u 2 L 1 () whose distributional derivative Du is represented by a bounded R N -valued measure on. We denote by BV (; 1) the class of all u 2 BV () which take values 1 only. For every function u on, Su is the set of all essential singularities, that is, the points of where u has no approximate limit; if u 2 BV () the set Su is rectiable, and this means that it can be covered up to an H N?1 -negligible subset by countably many hypersurfaces of class C 1. The essential boundary of a set E R N is the set @ E of all points in where E has neither density 1 nor density 0. A set E has nite perimeter in if its characteristic function 1 E belongs to BV (), or equivalently, if H N?1 (@ E \ ) is nite; in this case @ E is rectiable, and we may endow it with a measure theoretic normal E (dened up to H N?1 -negligible subsets) so that the measure derivative D1 E is represented as D1 E (B) = @ E\B E dh N?1 for every B. A function u :! 1 belongs to BV (; 1) if and only if fu = +1g (or fu =?1g as well) has nite perimeter in. In this case Su agrees with the intersection of the essential boundary of fu = +1g with, and the previous formula becomes Du(B) := 2 Su\B u dh N?1 for every B, (1:7) where u is a suitable normal eld to Su. We claim that Su is the interface between the phases fu = +1g and fu =?1g in the sense that it contains every point where both sets have density dierent than 0. For further results and details about BV functions and nite perimeter sets we refer the reader to [EG], chapter 5. 1.2. Hypotheses on J and W Unless dierently stated, the interaction potential J and the double-well potential W which appear in (1:1) satisfy the following assumptions: (i) J : R N! [0; +1) is an even function (i.e., J(h) = J(?h)) in L 1 (R N ) and satises R N J(h) jhj dh < 1 : (1:8) (ii) W : R! [0; +1) is a continuous function which vanishes at 1 only and has at least linear growth at innity (cf. the proof of Lemma 1.11). 1.3. The optimal prole problem and the denition of We rst dene the auxiliary unscaled functional F by F(u; A) := 1 4 x2a; h2r N J(h)? u(x + h)? u(x) 2 dx dh +? W u(x) dx (1:9) for every set A R N and every u : R N! R. Hence F(u; A) = F 1 (u; A) + 1 (u; A; R N n A). 4 x2a

We x now a unit vector e 2 R N and we denote by M the orthogonal complement of e. Hence every x 2 R N can be written as x = y + x e e where y is the projection of x on M and x e := hx; ei. We denote by C e the class of all (N? 1)-dimensional cubes centered at 0 which lie on M; for every C 2 C e, T C is the stripe T C := y +te : y 2 C; t 2 R, while Q C is the N-dimensional cube centered at 0 such that Q C \ M = C. Figure 1: the sets C, Q C and T C. A function u : R N! R is called C-periodic if u(x + re i ) = u(x) for every x and every i = 1; : : : ; N? 1, where r is the side length of C and e 1 ; : : : ; e N?1 are its axes. We denote by X(C) the class of all functions u : R N! [?1; 1] which are C-periodic and satisfy and nally we set lim u(x) = +1 and lim u(x) =?1 ; (1:10) x e!+1 x e!?1 (e) := inf jcj?1 F(u; C M = e T C ) : C 2 C e ; u 2 X(C) : (1:11) Q C The minimum problem (1:11) is called the optimal prole problem associated with the direction e, and a solution is called an optimal prole for transition T C I R e C in direction e. In [AB] we proved that the minimum in (1:11) is attained, and there exists at least one minimizer u which depends only on the variable x e, and more precisely u(x) = (x e ) where : R! [?1; 1] is the optimal prole associated with a certain one-dimensional functional F e. For the rest of this section is a xed regular open subset of R N. Theorem 1.4. Under the previous assumptions the following three statements hold: (i) Compactness: let be given sequences ( n ) and (u n ) L 1 () such that n! 0, and F n (u n ; ) is uniformly bounded; then the sequence (u n ) is relatively compact in L 1 () and each of its cluster points belongs to BV (; 1). (ii) Lower bound inequality: for every u 2 BV (; 1) and every sequence (u ) such that u! u in L 1 (), we have lim inf F (u ; ) F (u) ;!0 (iii) Upper bound inequality: u! u in L 1 () and for every u 2 BV (; 1) there exists a sequence (u ) such that lim sup F (u ; ) F (u) :!0 Remark 1.5. Statements (ii) and (iii) of Theorem 1.4 can be rephrased by saying that the functionals F (; ), or in short F,?-converge in the space L 1 () to the functional F given by (1:2) for all functions u 2 BV (; 1) and extended to +1 in L 1 () n BV (; 1). 5

For the general theory of?-convergence we refer the reader to [DM]; for the applications of?-convergence to phase transition problems we refer to the early paper of Modica [Mo], and to [Al] for a review of some of these results and related mathematical issues. Remark 1.6. As every?-limit is lower semicontinuous, we infer from the previous remark that the functional F given in (1:2) is weakly* lower semicontinuous and coercive on BV (; 1). The coercivity of F implies that the inmum of (e) over all unit vectors e 2 R N is strictly positive, while the semicontinuity implies that the 1-homogeneous extension of the function to R N, namely the function x 7! jxj? x jxj, is convex (see for instance [AmB], Theorem 3.1). Notice that it is not immediate to recover this convexity result directly from the denition of in (1:11). Remark 1.7. Statement (iii) of Theorem 1.4 can be rened by choosing the approximating u for every (we will not prove this renement of statement R R sequence (u ) so that u = (iii); in fact one has to slightly modify the construction of the approximating sequence (u ) given in Theorem 5.2). This way we can t with a prescribed volume constraint: given c 2 (?jj; jj), then the functionals R F?-converge to F also on the class Y c of all u 2 L 1 () which satisfy the volume constraint u = c. A sequence (v ) in Y c is called a minimizing sequence if v minimizes F (; ) in Y c for every > 0, and is called a quasi-minimizing sequence if F (v ; ) = inf F (u; ) : u 2 Y c + o(1). Using the semicontinuity result given in [AB], Theorem 4.7, and the truncation argument given in Lemma 1.11 below, we can prove that a minimizer of F (; ) in Y c exists provided that W is of class C 2 and W (t)?d for every t 2 [?1; 1], where d is dened by d := ess inf x2 Notice that d tends to 1 4 kjk 1 as! 0. 1 2 J (x 0? x) dx 0 : Now by a well-known property of?-convergence and by statement (i) of Theorem 1.4 we infer the following: Corollary 1.8. Let (v ) be a minimizing or a quasi-minimizing sequence for F on Y c. Then (v ) is relatively compact in L R 1 (), and every cluster point v minimizes F among all functions u 2 BV (; 1) which satisfy u = c. Equivalently, the set E := fv = 1g solves the minimum problem n o min ( E ) dh N?1 : E has nite perimeter in and jej = 1(c + jj) : 2 @ E 1.9. Outline of the proof of Theorem 1.4 for N = 1 In order to explain the idea of the proof of Theorem 1.4 and the connection with the optimal prole problem, now we briey sketch the proof of statement (ii) and (iii) for the one-dimensional case (the proof of statement (i) being slightly more delicate). In this case becomes the inmum of F 1 (; R) over the class X of all u : R! [?1; 1] which converge to +1 at +1 and to?1 at?1 (cf. (1:11)). We assume for simplicity that is the interval (?1; 1), and that u(x) =?1 for x < 0, u(x) = +1 for x 0. Then Su = f0g, and H 0 (Su) = ; a standard localization argument can be used to prove the result in the general case (cf. [Al], section 3a). We rst remark that the functionals F satisfy the following rescaling property: given > 0 and u : R! R we set u (x) := u(x), and then a direct computation gives F (u; R) = F 1 (u ; R) : (1:12) 6

Let us consider now the lower bound inequality. First we reduce to a sequence (u ) which converges to u in L 1 () and satises ju j 1; then we extend each u to the rest of R by setting u (x) :=?1 for x?1, u (x) := 1 for x 1. The key point of the proof is to show that F (u ; ) ' F (u ; R) as! 0. (1:13) By identity (1:5), (1:13) can be written in term of the locality defect, and more precisely it reduces to (u ; ; R n ) = o(1); notice that in general this equality may be false, but using the decay estimates for the locality defect given in section 2 we can prove that it is true if we replace with another interval, which may be chosen arbitrarily close to. By (1:12) and the denition of we get F (u ; R) = F 1 (u ; R), and then (1:13) yields lim inf!0 F (u ; ) : The proof of the upper bound inequality is even more simple: we take an optimal prole (i.e., a solution of the minimum problem which denes ) and we set u (x) := (x=) for every > 0. Then u (x) converge to u(x) for every x 6= 0, and (1:12) yields F (u ; ) F (u ; R) = F 1 (; R) = : Remark 1.10. It is clear from this brief sketch that the shape of the optimal prole plays no r^ole in the proof of Theorem 1.4, nor it does the fact that the minimum in (1:11) is attained: in case no optimal proles were available, it would suces to replace with functions in X which \almost minimize F 1 (; R). This could be indeed the case when one considers the vectorial version of this problem (see paragraph 1.12) Nevertheless the existence of the optimal prole has a deeper meaning than it appears above. Indeed if (v ) is a sequence of minimizers of F which converges to some v 2 BV (; 1), then we would expect that if we blow-up the functions v at some xed singular point x of v by taking the functions (x) := v ((x?x)), then \resembles more and more an optimal prole. In other words we expect the optimal proles to be the asymptotic shapes of the minimizers v about the discontinuity points of v. Yet a precise statement in this direction is far beyond the purposes of this paper. Concerning proofs, statements (i), (ii), and (iii) of Theorem 1.4 will be proved in sections 3, 4, and 5 respectively, while section 2 is devoted to the asymptotic estimates R for the locality defect. In section 6 we study the relation between the Dirichlet integral jruj2 and the interaction energy G (u) := J (x 0 u(x 0 )? u(x) 2dx? x) 0 dx : Warning: throughout the proof of Theorem 1.4, that is, in all sections from 2 to 5, we will always restrict ourselves to functions which take values in [?1; 1]. We are allowed to do this in view of the following lemma: Lemma 1.11. For every function u :! R, let T u denote the truncated function T u(x) :=? u(x) ^ 1) _?1. Then F (u; ) F (T u; ) for every > 0, and for every sequence (u ) such that F (u ; ) is bounded in there holds ku? T u k 1! 0 as! 0. Proof. The inequality F (u; ) F (T u; ) is immediate. Let now be given a sequence (u ) such that F (u ; ) C for every. Since W is strictly positive and continuous out of 1, and has growth at least linear at innity (see paragraph 1.2), for every > 0 we may nd a > 0, M > 0 and b > 0 so that 7

W (t) a when 1 + jtj M and W (t) bjtj when M jtj. Then we dene A and B as the sets of all x 2 where u (x) satises respectively 1 + u (x) M and M u (x). Hence ku? T u k 1 jj + MjA j + B ju j jj + M a + 1 b W (u ) : Since R W (u ) C, passing to the limit as! 0 we obtain that lim sup ku? T u k 1 jj, and since can be taken arbitrarily small the proof is complete. We conclude this section with a short overlook of the possible generalizations of Theorem 1.4 and some open problems. 1.12. The multi-phase model In order to describe a multi-phase system one may postulate a free energy of the form (1:1) where u is a vector density function on a domain of R N taking values in R m, W : R m! [0; 1) is a continuous function which vanishes at k + 1 anely independent wells f 0 ; : : : ; k g (and therefore k m), and J is the usual interaction potential. Theorem 1.4? holds provided we make the following modications: BV (; 1) is replaced by the class BV ; f i g of all functions u 2 BV (; R m ) which takes values in f 0 ; : : : ; k g only, and the functional F is now dened by F (u) := X i<j S ij ij ( ij ) dh N?1 ; (1:14) where S ij is the interface which separates the phases fu = i g and fu = j g, and precisely S ij := @ fu = i g \ @ fu = j g \ (recall that both phases have nite perimeter in ), and ij is the measure theoretic normal to S ij. For every unit vector e the value ij (e) is dened by the following version of the optimal prole problem: ij (e) := inf jcj?1 F(u; T C ) : C 2 C e ; u 2 X ij (C) ; (1:15) here we follow the notation of paragraph 1.3 and X ij (C) is the class of all functions u : R N! R m which are C-periodic and satisfy the boundary condition lim u(x) = j and lim u(x) = i : x e!+1 x e!?1 This vectorial generalization of Theorem 1.4 can be proved by adapting the proof for the? scalar case given below, and using a suitable approximation result for the functions in BV ; f i g (cf. the approach in [Ba] for the vectorial version of the Modica-Mortola theorem). Notice that in this case it is not known whether the optimal prole problem (1:15) admits a solution or not (cf. [AB], section 4b). 1.13. The optimal assumptions on J As already remarked, in the current approach the ferromagnetic condition J 0 cannot be removed; more precisely it plays an essential r^ole in the proof of statement (i) of Theorem 1.4, and in particular in the rst step of the proof of Theorem 3.1 (on the other hand the proofs of statements (ii) and (iii) do not require the non-negativity of J). Yet it would be quite interesting to understand the asymptotic behaviour of the functionals F when J is allowed to take also negative values. As far as we know in that case it may well happen that dierent scalings should be considered for the functional F, and that the?-limit has a completely dierent form. 8

About the growth assumptions on J, we can replace the hypotheses in paragraph 1.3, namely J 2 L 1 (R N ) and (1:8), with the following more general ones (cf. [AB], section 4c): J is even, non-negative, and satises R N J(h)? jhj ^ jhj 2 dh < +1 : (1:16) We remark that the proof of Theorem 1.4 needs no modications at all if J does not belong to L 1 (R N ) but still veries (1:8), while some additional cares have to be taken in the fully general case, and more precisely in the proof of stament (iii) (see in particular the third step in the proof of Theorem 5.2 and the decay of the locality defect in Lemma 2.7). Indeed statements (i) and (ii) can always be recovered from the usual version of Theorem 1.4 by approximating J with an increasing sequence of potentials which satisfy the assumptions in paragraph 1.2. Finally we notice that if (1:16) does not hold, then the value of (e) as given by the optimal prole problem (1:11) is always equal to +1 (cf. [AB], Theorem 4.6). This probably means that a dierent scaling should be considered in the denition of the functionals F. For instance, if N = 1 and J(h) = 1=h 2 the \right scaling is given by u(x0 )? u(x) x 0? x 2 dx 0 dx + e 1= W? u(x) dx ; or equivalently by multiplying the functionals F dened in (1:1) by an innitesimal factor of order j log j?1. In this case we obtain again a?-limit of the form (1:2) (see for instance [ABS]). However no general result is available when J does not veries (1:16). Acknowledgements: the authors are deeply indebted with Errico Presutti for many valuable discussions and for his constant support. They would also like to thank Marzio Cassandro. This paper has been completed while the rst author is visiting for one year the Max Planck Institut for Mathematics in the Sciences in Leipzig. 2. Decay estimates for the locality defect. In this section we study the asymptotic behaviour as tends to 0 of the locality defect. Roughly speaking, the goal is to show that the limit of (u ; A; A 0 ) is determined only by the asymptotic behaviour of the sequence u close to the intersection of the boundaries of A and A 0. The main result of this section is Theorem 2.8. We rst need to x some additional notation. We dene the auxiliary potential ^J by ^J(h) := 1 0 J h h t dt for every t t N h 2 R N. (2:1) It follows immediately from the denition that ^J is even, non-negative, and satises k ^Jk1 = R N J(h) jhj dh : (2:2) Denition 2.1. Throughout this section always denotes a subset of a Lipschitz hypersurface in R N, and is endowed with the Hausdor measure H N?1 ; we often omit any explicit mention to this measure. 9

Let A be a set of positive measure in R N, and take a sequence (u n ) of functions from A into [?1; 1], and a sequence ( n ) of positive real numbers which tends to 0. We say that the n -traces of u n (relative to A) converge on to v :! [?1; 1] when lim h n!1 y2 fh: y+ nh2ag ^J(h) un (y + n h)? v(y) dh i dy = 0: (2:3) Remark 2.2. Notice that we make no assumption on the relative position of A and ; in particular they may be even distant. Notice moreover that the notion of \convergence of the n -traces is introduced without dening what the n -trace of a function is, and in fact there is no such notion. This is due to the fact that for functions in the domain of F, the trace on an (N? 1)-dimensional manifold cannot be dened (while it is dened for functions in the domain of the?-limit, that is, for BV functions). In view of the denition of the locality defect, it would make more sense to replace the term un (y + n h)? v(y) in (2:3) with its square. But since we restrict ourselves to functions which take values in [?1; 1], the limit in (2:3) is independent of the exponent of un (y + n h)? v(y), and we chose 1 because it simplies many of the following proofs. Remark 2.3. We can dene the upper ^J-density of A at the point x 2 R N as the upper limit lim sup ^J(h) dh ;!0 fh: x+h2ag and the lower ^J-density as the corresponding lower limit; notice that such densities are local, that is, they do not depend on the behaviour of A out of any open neighborhood of x. The function v which satises (2:3) is uniquely determined for (H N?1 -) almost every point of where A has positive J-upper density. If (2:3) holds for some set A, then it is veried by every A 0 included in A. Moreover if has nite measure then (2:3) is also veried by every A 0 such that A 0 n A has upper ^J-density 0 at almost every point of. In particular if are given sets A and A 0 such that the symmetric dierence A4A 0 has upper ^J-density 0 at almost every point of, then A satises (2:3) if and only A 0 does. Remark 2.4. when Condition (2:3) is not easy to verify. If has nite measure then (2:3) holds lim u n(y + n h) = v(y) for a.e. y 2 and a.e. h 2 A. (2:4) n!1 Condition (2:4) holds for instance when u n converge locally uniformly on some open neighborhood of to a function which, at every point of, is continuous and agrees with v. Assume now that the functions u n converge to u in L 1 (A). Unfortunately this is not enough to deduce that the n -traces of u n converge to u on every Lipschitz hypersurface R N, yet this holds for \most. More precisely, we have the following proposition: Proposition 2.5. Take A, ( n ) and (u n ) as in Denition 2.1; let g : A! R be a Lipschitz function, and denote by t the t-level set of g for every t 2 R. If u n! u in L 1 (A) then, possibly passing to a subsequence, the n -traces of u n (relative to A) converge to u on t for a.e. t 2 R. (Since g admits a Lipschitz extension to R N, t is a subset of an oriented closed Lipschitz hypersurface in R N for almost every t 2 R.) 10

Proof. To simplify the notation we write ; u instead of n ; u n, we assume that g is 1-Lipschitz and A = R N (the general case follows in the same way). For every > 0, x 2 R N and t 2 R we set (x) := ^J(h) u (x + h)? u(x) dh and g (t) := R N By the coarea formula for Lipschitz functions (see [EG], section 3.3) we get g (t) dt = R = RN (x) rg(x) dx R N (x) dx t (x) dx : (2:5) N^J(h) u (x + h)? u(x) dx dh R N R h RNRN^J(h) u (x + h)? u(x + h) + u(x i + h)? u(x) dx dh ^J(h) ku? uk 1 + k h u? uk 1 dh ; (2:6) R N where h u(x) := u(x + h). Now ku? uk 1 tends to 0 by assumption and k h u? uk 1 tends to 0 as! 0 for every h, and since ^J is summable (cf. (2:2)) we can apply the dominated convergence theorem to the integrals in line (2:6), and we get lim g (t) dt = 0 :!0 R Hence the functions g converge to 0 in L 1 (R), and passing to a subsequence we may assume that they also converge pointwise to 0 for a.e. t 2 R. Since g (t) is equal to the double integral in (2:3) (with v replaced by u), the proof is complete. Denition 2.6. Let be given A; A 0 R N. We say that the set divides A and A 0 when for every x 2 A, x 0 2 A 0 the segment [x; x 0 ] intersects. We say that strongly divides A and A 0 when is the (Lipschitz) boundary of some open set such that A and A 0 R N n. Now we can state and prove the rst decay estimate for the locality defect. Let be given disjoint sets A and A 0 in R N which are divided by, then take positive numbers n! 0 and functions u n : A [ A 0! [?1; 1] and v; v 0 :! [?1; 1]. Lemma 2.7. Under the above stated hypotheses, if the n -traces of u n relative to A and A 0 converge on to v and v 0 respectively, then lim sup n (u n ; A; A 0 ) 1 n!1 2 k ^Jk1 v(y)? v 0 (y) dy : (2:7) Proof. To simplify the notation we write, u and instead of n, u n, n. By the denition of, and recalling that ju j 1, we obtain (u ; A; A 0 ) 1 2 R N J(h) h A h u (x + h)? u (x) dx {z } I (h) 11 i dh ; (2:8)

where A h is the set of all x such that x 2 A and x + h 2 A 0. Let us consider for the moment the integral I (h) dened in (2:8): for every x in the integration domain A h, the segment [x; x + h] intersects, and then we can write x as x = y? th for some y 2 and t 2 [0; 1]. Figure 2: the set A h for given > 0 and h 2 R N. Since the Jacobian determinant of the map which takes (y; t) 2 [0; 1] into y? th does not exceed jhj, by applying the change of variable x = y? th we get I (h) jhj A where S hy is the set of all t 2 [0; 1] such that y? th 2 A and y + (1? t)h 2 A 0. Hence (2:8) yields x εh (u ; A; A 0 ) 1 2 h2r N ; y2 hshy u (y + (1? t)h)? u (y? th) dt i h J(h) jhj S hy A εh y Σ x'=x+εh dy ; u (y + (1? t)h)? u (y? th) dt i A' dy dh : (2:9) Now by the triangle inequality we can estimate u (y + (1? t)h)? u (y? th) by the sum of the following three terms: v(y)? v 0 (y) + u (y? th)? v 0 (y) + u (y + (1? t)h)? v(y) : Accordingly we estimate the double integral at the right hand side of (2:9) by the sum of the corresponding double integrals I 1, I 2 and I 3, that is, We recall now that js hy j 1 for every h and every y, and then I 1 : = 1 2 1 2 (u ; A; A 0 ) I 1 + I 2 + I 3 : (2:10) h2r N ; y2 h J(h) jhj h S hy R N J(h) jhj dh v(y)? v 0 (y) dt i dy dh ih i v(y)? v 0 (y) dy : (2:11) Since the rst integral in line (2:11) is equal to k ^Jk1 (see (2:2)), inequality (2:7) will follow from (2:10) once we have proved that I 2 and I 3 vanish as! 0. Let us consider I 2 : I 2 : = 1 2 h2r N ; y2 J(h) jhj hshy u (y? th)? v(y) dt i dy dh 12

1 (1) 2 (2) 1 2 h 0 2R N ; y2 h h S hy J fh 0 : y+h 0 2Ag h 0 t h 0 u (y + h 0 )? v(y) dt t ^J(h 0 ) u (y + h 0 )? v(y)j dh 0i dy : i t N dy dh 0 Hence I 2 vanishes as! 0 because the -traces of u relative to A converge to v on. In a similar way one can prove that I 3 vanishes as! 0. Now we can state the main result of this section. Let be given disjoint sets A; A 0 R N, and such that one of the following holds: (a) the sets A and A 0 are divided by (cf. Denition 2.6); (b) the sets A and A 0 are strongly divided by a Lipschitz boundary S with nite measure and = @A \ @A 0 ; (c) either A or A 0 is a bounded set with Lipschitz boundary and = @A \ @A 0. Take then positive numbers n! 0 and functions u n : A [ A 0! [?1; 1]. Theorem 2.8. Under the above stated hypotheses we have lim sup n!1 n (u n ; A; A 0 ) k ^Jk1 H N?1 () : (2:12) Moreover if the n -traces of u n relative to A and A 0 converge on respectively to v and v 0, then lim sup n (u n ; A; A 0 ) 1 n!1 2 k ^Jk1 v(y)? v 0 (y) dy : (2:13) Proof. Notice that (2:12) follows by applying (2:13) to the functions u n which are equal to 1 on A and to?1 on A 0 (with v = 1 and v 0 =?1) and then using the obvious inequality n (u n ; A; A 0 ) n (u n ; A; A 0 ). Let us prove (2:13). When (a) holds it is enough to apply Lemma 2.7, while (c) clearly implies (b). Assume that (b) holds. First of all we notice that in this case we can always modify the boundary S so that S \ @A = S \ @A 0 =. Now we extend v and v 0 to 0 in S n, and then the n -traces of u n relative to A and A 0 converge on S to v and v 0 respectively (use Remark 2.3, recalling that both A and A 0 have upper J-density 0 at every point of S n ). Now it is enough to apply Lemma 2.7 with S instead of. 3. Proof of the compactness result The following theorem implies statement (i) of Theorem 1.4, and shows that the domain of the?-limit of the functionals F is included in BV (; 1). Theorem 3.1. Let be a regular open set and let be given sequences ( n ) and (u n ) such that n! 0, u n :! [?1; 1], and F n (u n ; ) is bounded. Then the sequence (u n ) is relatively compact in L 1 () and each of its cluster points belongs to BV (; 1). (1) Apply the change of variable h =?h 0 =t. (2) Recall (2:1) and that js hy j 1. 13

Proof. In order to simplify the notation we replace as usual n, u n, and F n with, u, and F. We need the following inequality, which may be proved by a direct computation: for every non-negative g 2 L 1 (R N ) and every u : R N! R there holds R N R N (g g)(y) u(x + y)? u(x) dy dx 2kgk1 R N R N g(y) u(x + y)? u(x) dy dx : (3:1) The proof of the theorem is now divided into two steps. Step 1. We rst prove the thesis under the assumption that each u takes values 1 only. We extend each function u to 1 in R N n, and then we observe that R N R N J (y) u (x + y)? u (x) dy dx = O() : (3:2) Indeed the assumption u = 1 implies u (x 0 )? u (x) = 1 denition of F we obtain 1 J (x R N R N 0? x) u (x 0 )? u (x) dx 0 dx = 2? u (x 0 )? u (x) 2, and then by the = 2F (u ; R N ) = 2F (u ; ) + 4 (u ; ; R N n ) : We apply inequality (2:12) with A = and A 0 = R N n to show that (u ; ; R N n ) is uniformly bounded in (recall that we are considering only a subsequence n which converges to 0), while F (u ; ) is uniformly bounded by assumption. Hence (3:2) is proved. Now we combine inequality (3:1) with g := J and inequality (3:2), and we obtain R N R N (J J )(y) u (x + y)? u (x) dy dx = O() : (3:3) Since J J is a non-negative continuous function, we may nd a non-negative smooth function ' (not identically 0) with compact support such that ' J J and jr'j J J : (3:4) We set c := R R N '(y) dy and for every y 2 R N and every > 0 we dene ' (y) := 1 c N '(y=) and w (y) := ' u (y) : (3:5) The functions ' are smooth and non-negative, have integral equal to 1, and converge to the Dirac mass centered at 0 as! 0. We claim that the sequence (w ) is asymptotically equivalent to (u ) in L 1 (R N ), and that the gradients rw are uniformly bounded in L 1 (R N ). Once this claim is proved we could infer that the sequence (w ) is relatively compact in L 1 () and each of its cluster points belongs to BV (; 1), and the same holds for the sequence (u ). Now it remains to prove the claim. We have R N jw? u j dx = RN? ' (y) u (x + y)? u (x) dy dx R N ' (y) u (x + y)? u (x) dy dx R N R N (J J )(y) u (x + y)? u (x) dy dx = (4) O() : 1 (3) c R N R N (3) By (3:4) and (3:5) we obtain ' 1 c J J. (4) Apply estimate (3:3). 14

Moreover jrw j dx = R N = (5) RN r' (y) u (x + y) dy dx R N? r' (y) u (x + y)? u (x) dy dx R N R N r' (y) u (x + y)? u (x) dy dx R N R N (J J )(y) u (x + y)? u (x) dy dx = (7) O(1) ; 1 (6) c R N R N and the claim is proved. Step 2. We consider now the general case. For every s 2 R we set and then we dene T (s) :=?1 if s < 0, +1 if s 0, (3:6) v := T (u ) : (3:7) The functions v takes values 1 only, and we claim that the sequence (v ) is asymptotically equivalent to (u ) in L 1 () and that F (v ; ) is uniformly bounded. Once proved this claim the thesis will follow from Step 1. Take so that 0 < < 1, and let K be the set of all x 2 such that u (x) 2 [?1+; 1?]. Then ju? v j in n K, and we deduce ju? v j dx jj + K? ju j + jv j dx jj + 2jK j : (3:8) Since > 0 and W is zero only at 1, there exists a positive constant (which depends on ) such that W (t) for every t 2 [?1 + ; 1? ]. Hence jk j 1 Inequalities (3:8) and (3:9) imply K W? u (x) dx F (u ; ) = O() : (3:9) lim sup!0 ju? v j dx jj : As is arbitrary, the sequences (u ) and (v ) are asymptotically equivalent in L 1 (). R It remains to prove that F (v ; ) is uniformly bounded in. Since W (v ) dy = 0, we have only to estimate the rst integral in the denition of F. Given s 1 ; s 2 2 [?1; 1] we have that either js 1 j 1=2 or T (s1 )? T (s 2 ) 4js1? s 2 j. (5) Recall that R R N r' (y) dy = 0 because ' has compact support. (6) By (3:4) and (3:5) we obtain jr' j 1 c J J. (7) Apply estimate (3:3). 15

Hence, if we denote by H the set of all x 2 such that u (x) 1=2, we deduce F (v ; ) = 1 4 J (x 0? x)? T u (x 0 )? T u (x) 2 dx 0 dx 4? J (x 0? x) u (x 0 )? u (x) 2 dx 0 dx + 1 J (x 0? x) dx 0 dx H 16 F (u ; ) + 1 kjk 1jH j : (3:10) By the properties of W there exists a positive constant such that W (t) for every t such that jtj 1=2, and reasoning as in (3:9) we get jh j = O(); together with (3:10) this proves that F (v ; ) is uniformly bounded in. 4. Proof of the lower bound inequality In this section we prove statement (ii) of Theorem 1.4. We begin with some notation. For every > 0, A R N and u : R N! [?1; 1] we dene the rescaling of the functional F given in (1:9) by F (u; A) := 1 4 x2a; h2r N J (h)? u(x + h)? u(x) 2 dx dh + 1 Recalling the denitions of F and we obtain: x2a W? u(x) dx : (4:1) F (u; A) = F (u; A) + (u; A; R N n A) : (4:2) Let be given now a function u dened on (a subset of) R N, a point x 2 R N and a positive number r. We dene the blow-up of u centered at x with scaling factor r the function R x;r u given by (R x;r u)(x) := u(x + rx) ; (4:3) when x = 0 we write R r u instead of R 0;r u. For every set A R N we set, as usual, x + ra := fx + rx : x 2 Ag, and then we easily obtain the following scaling identities: F (u; x + ra) = r N?1 F =r (R x;r u; A) (4:4) F (u; x + ra) = r N?1 F =r (R x;r u; A) : (4:5) In the proof we also make use of the following well-known results about the blow-up of nite perimeter sets and measures: 4.1. Some blow-up results Let S be a rectiable set in R N with normal vector eld ; let be the restriction of the Hausdor measure H N?1 to the set S, that is, := H N?1 S, and let be a nite measure on R N. Then for H N?1 -a.e. x 2 S the density of with respect to at x is given by the following limit:? d x + rq (x) = lim (4:6) d r!0 r N?1 where Q is any unit cube centered at 0 such that (x) is one of its axes. 16

Let u be a xed function in BV (; 1). For every x 2 Su we denote by v x : R N! 1 the step function +1 if hx; u (x)i 0, v x (x) := (4:7)?1 if hx; u (x)i < 0. Then for H N?1 -a.e. x 2 Su, and more precisely for all x 2 Su such that the density of the measure Du with respect to jduj exists and is equal to u (x), there holds R x;r u?! v x in L 1 loc (RN ) as r! 0 (4:8) (if u is not dened on the whole of R N we take an arbitrary extension). 4.2. Proof of statement (ii) of Theorem 1.4 We can now begin the proof of statement (ii) of Theorem 1.4. We assume therefore that is given a sequence (u ) which converges to u 2 BV (; 1) in L 1 (); we have to prove that lim inf!0 F (u ; ) Su ( u ) dh N?1 : (4:9) In the following u and u are xed. We shall often extract from all positive a subsequence ( n ) which converges to zero; to simplify the notation we shall keep writing, F, and u instead of n, F n, u n. First of all we notice that it is enough to prove inequality (4:9) when the lower limit at the left hand side is nite and then, passing to a subsequence we may assume as well that it is a limit. Now we follow the approach of [FM]; the main feature of this method consists in the reduction of the lower bound inequality (4:9) to a density estimate (see (4:13)) which has to be veried point by point. What follows, up to equation (4:18), is a straightforward adaptation of this general method (see also [BF], [BFM]). For every > 0 we dene the energy density associated with u at the point x 2 as g (x) := 1 4 and then we consider the corresponding energy distribution J (x 0? x)? u (x 0 )? u (x) 2 dx 0 + 1 W? u (x) ; (4:10) := g L N : (4:11) Thus the total variation k k of the measure (on ) is equal to F (u ; ), and since F (u ; ) is equibounded with respect to, possibly passing to a subsequence we can assume that there exists a nite positive measure on such that * weakly* on as! 0. Since F (u ; ) = k k and lim inf k k kk, inequality (4:9) is implied by the following:!0 kk Su\ ( u ) dh N?1 : (4:12) In fact, we prove a stronger result: the density of with respect to := H N?1 than or equal to ( u ) at H N?1 -a.e. point of Su, that is Su is greater d d (x)? u (x) for H N?1 -a.e. x 2 Su. (4:13) 17

More precisely, we have the following lemma. Lemma 4.3. With the previous notation, inequality (4:13) holds for every x 2 Su which veries (4:6) and (4:8). Proof. We x such a point x 2 Su, and we denote by the vector u (x) and by v the step function v x dened in (4:7). Following the notation of paragraph 1.3 we x an (N? 1)-dimensional unit cube C 2 C, and we take Q = Q C and T = T C accordingly. As the measures weak* converge? to on as! 0, we have that (A)! (A) for every set A such that (@A) = 0. Since x + r(@q) = 0 for all positive r up to an exceptional countable set N, we deduce that (x + rq)! (x + rq) for every positive r =2 N. Therefore, recalling (4:6) we write lim r!0 r =2N lim!0 (x + rq) r N?1 (x + rq) = lim r!0 r N?1 r =2N Since u! u in L 1 () by assumption and (4:8) holds, we also have lim r!0 = d (x) : (4:14) d? lim!0 R x;r u = lim r!0 R x;r u = v in L 1 (Q). (4:15) Therefore by a diagonal argument we may choose sequences (r n ) and ( n ) so that lim r n = lim n!1 n!1 lim n!1 n (x + r n Q) r N?1 n? n rn = 0 ; (4:16) = d (x) ; (4:17) d lim R x;r n!1 n u n = v in L 1 (Q), (4:18) and then we set n := n =r n, v n := R x;rn u n. To simplify the notation in the following we write,, r, u and v instead of n, n, r n, u n and v n respectively. >From the scaling identity (4:4) and the denition of we infer (x + rq) r N?1 F (u ; x + rq) r N?1 = F (v ; Q) : (4:19) Keeping in mind (4:17) and (4:19), we can try to prove (4:13) by establishing a precise relation between F (v ; Q) and () (see paragraph 1.3). One possibility is the following: we extend v to the strip T by setting v := v in T n Q, and then we take the C-periodic extension in the rest of R N. Now, by the scaling identity (4:5) we know that F (v ; T ) () ; and then it would remain to prove that the dierence between F (v ; T ) and F (v ; Q) vanishes as! 0; this dierence can be written as (cf. (4:20) below) F (v ; T )? F (v ; Q) = (v ; T; R N n T ) + 2 (v ; Q; T n Q) ; but unfortunately we cannot use Theorem 2.8 to show that it vanishes as! 0 because we have no information about the convergence of the -traces of v on the boundaries @Q and @T. We overcome this diculty as follows: as v! v in L 1 (Q), Theorem 2.8 shows that for a.e. t 2 (0; 1) the -traces of v converge to v on the boundary t of the cube tq (notice that each t is the t-level set of the Lipschitz function g(x) := dist(x; @Q)). 18

We x for the moment such a t, and we dene ~v on the stripe tt as ~v (x) := 8 < : v (x) v(x) if x 2 tq, if x 2 tt n tq, and then we take the tc-periodic extension in the rest of R N. Hence ~v belongs to X(tC) (cf. paragraph 1.3), and since ~v = v in tq F (v ; Q) F (v ; tq) = F (~v ; tq) = = F (~v ; tt )? 2 (~v ; tq; tt n tq) = F (~v ; tt )? (~v ; tt; R N n tt ) {z } L 1?2 (~v ; tq; tt n tq) {z } : (4:20) L 2 Now we claim that both locality defects L 1 and L 2 vanish as! 0; once this is proved we could deduce from the previous formula that lim sup!0 F (v ; Q) lim sup!0 F (~v ; tt ) : (4:21) Let us consider rst L 2 : the sets tq and tt n tq are divided by the boundary t of tq, and by the choice of t the -trace of ~v relative to tq converge to v on t (recall that ~v = v on tq). On the other hand ~v = v in tt n tq, and then also the -trace relative to tt n tq converge to v on t. Hence Theorem 2.8 applies, and L 2 vanishes as! 0. In a similar way one can prove that also L 1 vanishes as! 0 (it is enough to verify that the -trace of ~v relative to R N converge to v on the boundary of tt ). Eventually we use the scaling identity (4:5) and the denition of () to get F (~v ; tt ) = N?1 F? R ~v ; t T t N?1 () ; (4:22) and putting together (4:17), (4:19), (4:21) and (4:22) we obtain d d (x) tn?1 () ; the proof of inequality (4:13) is thus completed by taking t arbitrarily close to 1. 5. Proof of the upper bound inequality Throughout this section is always a regular open set. Denition 5.1. A N-dimensional polyhedral set in R N is an open set E whose boundary is a Lipschitz manifold contained in the union of nitely many ane hyperplanes; the faces of E are the intersections of the boundary of E with each one of these hyperplanes, and an edge point of E is a point which belongs to at least two dierent faces (that is, a point where @E is not smooth). We denote by E the inner normal to @E (dened for all points in the boundary which are not edge points). A k-dimensional polyhedral set in R N is a polyhedral subset of a k-dimensional ane subspace of R N. A polyhedral set in is the intersection of a polyhedral set in R N with. 19

We say that u 2 BV (; 1) is a polyhedral function if there exists an N-dimensional polyhedral set E in R N such that @E is transversal to @ (that is, H N?1 (@E \ @) = 0) and u(x) = 1 for every x 2 \ E, u(x) =?1 for every x 2 n E. Theorem 5.2. Let u 2 BV (; 1) be a polyhedral function. Then there exists a sequence of functions (u ) dened on such that ju j 1 for every, u converge to u uniformly on every compact set K n Su, and lim sup F (u ; )!0 Su ( u ) dh N?1 : (5:1) Proof. Let us x some notation: E is the polyhedral set associated with u in Denition 5.1; we denote by S the set of all edge points of E which belongs to and by a general face of Su (that is, a face of E). Then S is a nite union of (N? 2)-dimensional polyhedral sets in, @E = Su, and we may choose the orientation of Su so that E = u (for every point in Su n S). Given open sets A 1 ; A 2 let A 1 t A 2 we denote the interior of A 1 [ A 2. We dene G as the class of all sets A such that (i) A is an N-dimensional polyhedral set in, and @A and Su are transversal (that is, H n?1 (Su \ @A) = 0); (ii) there exists a sequence of functions (u ) dened on A such that ju j 1 and u! u uniformly on every compact set K A n Su, (5:2) lim sup F (u ; A)!0 A\Su ( u ) dh N?1 : (5:3) The proof of Theorem 5.2 is achieved by showing that 2 G; this is a consequence of the following three statements: (a) if A is an N-dimensional polyhedral set in such that H N?1 (A \ Su) = 0, then A 2 G; (b) let be a face of Su and let be the projection map on the ane hyperplane which contains : if A is an N-dimensional polyhedral set in such that Su \ A = and (A) =, then A 2 G; (c) if A 1 ; A 2 belong to G and are disjoint, then A 1 t A 2 2 G. Step 1: proof of statement (a). In this case H N?1 (@A \ Su) = 0 and A \ Su = ; then u is constant (?1 or 1) in A, and it is enough to take u := u for every > 0. Step 2: proof of statement (b). Property (i) is immediate; let us prove (ii). We denote by e the (constant) inner normal to ; therefore lies on some ane hyperplane which is parallel to M; without loss of generality we may assume that lies exactly in M. Following the notation of paragraph 1.3, for every xed > 0, we can nd C 2 C e and w 2 X(C) such that jcj?1 F(w; T C ) (e) + ; (5:4) and then we dene u (x) := w(x=) for every x 2 R N. (5:5) Property (5:2) holds because w(x)! 1 as x e! 1 (see paragraph 1.3). We claim that? lim sup F (u ; A) H N?1 () (e) + : (5:6)!0 20

Without loss of generality, we may assume that C is a unit cube. In order to prove inequality (5:6), for every > 0 we cover with the closures of a nite number h = h() of pairwise disjoint copies of the (N? 1)-dimensional cube C, that is, we choose x 1 ; : : : ; x h 2 M so that h[? xi + C : i=1 Moreover, since is a polyhedral set in M, the previous covering can be chosen so that h N?1 = H N?1? [ (x i + C)?! H N?1 () as! 0. (5:7) Notice that, since is the projection of A on M, then A [ i (x i + T C ). Then by denition (4:1) we have F (u ; A) F? u ; [ i (x i + T C ) F? u ; [ i (x i + T C ) = hx i=1 F (u ; T C ) ; (5:8) where the last equality follows from the fact that F (u ; ) is translation invariant and additive. Applying now the scaling identity (4:5) with x = 0 and = r we get F (u ; T C ) = N?1 F(w; T C ), so that by (5:8) and (5:4) we deduce F (u ; A) h N?1? (e) + : Taking into account (5:7) we get (5:6). Since e agrees with u in = Su \ A, (5:3) follows from inequality (5:6) by a simple diagonal argument, and the proof of statement (b) is complete. Step 3: proof of statement (c). Given disjoint A 1 ; A 2 2 G, we set A := A 1 t A 2 and we take sequences (u 1 ), (u 2 ) which satisfy property (ii) for A 1 and A 2 respectively. Then we set u (x) := 8 < : u 1 (x) if x 2 A 1, u 2 (x) if x 2 A 2. One can check that properties (i) and (5:2) are satised, and that (5:3) reduces to lim (u ; A 1 ; A 2 ) = 0 :!0 Notice that by (5:2) the -traces of u i relative to A i converge to u on every Lipschitz hypersurface A i such that H N?1 ( \ Su) = 0 for i = 1; 2 (cf. Remark 2.4); in particular this holds true for = @A. Hence the previous identity follows from Theorem 2.8. Step 4: proof of Theorem 5.2. It may be veried that may be written as = ta i where the sets A i are nitely many, pairwise disjoint, and satisfy the hypothesis of statements (a) or (b): 21