• Aucun résultat trouvé

Approximation and relaxation of perimeter in the Wiener space

N/A
N/A
Protected

Academic year: 2021

Partager "Approximation and relaxation of perimeter in the Wiener space"

Copied!
31
0
0

Texte intégral

(1)

HAL Id: hal-00609031

https://hal.archives-ouvertes.fr/hal-00609031

Submitted on 17 Jul 2011

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires

Approximation and relaxation of perimeter in the Wiener space

Michael Goldman, Matteo Novaga

To cite this version:

Michael Goldman, Matteo Novaga. Approximation and relaxation of perimeter in the Wiener space.

Annales de l’Institut Henri Poincaré, 2012. �hal-00609031�

(2)

Approximation and relaxation of perimeter in the Wiener space

M. Goldman

M. Novaga

Abstract

We characterize the relaxation of the perimeter in an infinite dimensional Wiener space, with respect to the weak L

2

-topology. We also show that the rescaled Allen- Cahn functionals approximate this relaxed functional in the sense of Γ-convergence.

1 Introduction

Extending the variational methods and the geometric measure theory from the Euclidean to the Wiener space has recently attracted a lot of attention. In particular, the theory of functions of bounded variation in infinite dimensional spaces started with the works by Fukushima and Hino [21, 22]. Since then, the fine properties of BV functions and sets of finite perimeter have been investigated in [4, 5, 3, 1]. We point out that this theory is closely related to older works by M. Ledoux and P. Malliavin [26, 27].

In the Euclidean setting it is well-known that the perimeter can be approximated by means of more regular functionals of the form

Z ε

2 |∇ u |

2

+ W (u) ε

dx

when ε tends to zero, in the sense of Γ-convergence with respect to the strong L

1

-topology [29, 28]. An important ingradient in this proof is the compact embedding of BV in L

1

. A natural question is whether a similar approximation property holds in the infinite dimensional case. The main goal of this paper is answering to this question by computing the Γ-limit, as ε → 0, of the Allen-Cahn-type functionals (see Section 2 for precise definitions)

F

ε

(u) = Z

X

ε

2 |∇

H

u |

2H

+ W (u) ε

dγ.

CMAP, CNRS UMR 7641, Ecole Polytechnique, 91128 Palaiseau, France, email: gold- man@cmap.polytechnique.fr

Dipartimento di Matematica, Universit` a di Padova, via Trieste 63, 35121 Padova, Italy, email: no-

vaga@math.unipd.it

(3)

In the Wiener space there are two possible definitions of gradient, and consequently two different notions of Sobolev spaces, functions of bounded variation and perimeters [4, 1].

In one definition the compact embedding of BV

γ

(X) in L

1γ

(X) still holds [4, Th. 5.3] and the Γ-limit of F

ε

is, as expected, the perimeter up to a multiplicative constant. We do not reproduce here the proof of this fact, since it is very similar to the Euclidean one.

A more interesting situation arises when we consider the other definition of gradient, which gives rise to a more invariant notion of perimeter and is therefore commonly used in the literature [21, 22, 4]. In this case, the compact embedding of BV

γ

(X) in L

1γ

(X) does not hold anymore. In particular sequences with uniformly bounded F

ε

-energy are not generally compact in the (strong) L

1γ

-topology, even though they are bounded in L

2γ

(X), and hence compact with respect to the weak L

2γ

(X)-topology. This suggests that the right topology for considering the Γ-convergence should rather be the weak L

2γ

(X)-topology.

A major difference with the finite dimensional case is the fact that the perimeter function defined by

F(u) =

( P

γ

(E) if u = χ

E

+ ∞ otherwise

is no longer lower semicontinuous in this topology, and therefore cannot be the Γ-limit of the functionals F

ε

. The problem is that the sets of finite perimeter are not closed under weak convergence of the characteristic functions. However, it is possible to compute the relaxation F of F (Theorem 4.6), which reads:

F (u) =

 

 Z

X

q

U

2

(u) + | D

γ

u |

2

dγ if 0 ≤ u ≤ 1

+ ∞ otherwise.

Such functional is quite familiar to people studying log–Sobolev and isoperimetric in- equalities in Wiener spaces [6, 7, 10].

Our main result is to show that the Γ-limit of F

ε

, with respect to the weak L

2γ

(X)- topology, is a multiple of F (Theorem 5.3). The proof relies on the interplay between symmetrization, semicontinuity and isoperimetry.

The plan of the paper is the following. In Section 2 we recall some basic facts about Wiener spaces and functions of bounded variation. In Section 3 we give the main properties of the Ehrhard symmetrizations. We also prove a P´ olya-Szeg¨ o inequality and a Bernstein- type result in the Wiener space (Propositions 3.12 and 3.5), which we believe to be interesting in themselves. In Section 4, we use the Ehrhard symmetrization to compute the relaxation of the perimeter (Theorem 4.6). Finally, in Section 5 we compute the Γ-limit of the functionals F

ε

(Theorem 5.3) and discuss some consequences of this result.

Acknowledgements: The authors wish to thank Michele Miranda for valuable discus-

sions. The first author would like also to thank the Scuola Normale di Pisa for the kind

hospitality, and Luigi Ambrosio for the invitation and the interest in this work.

(4)

2 Wiener space and functions of bounded variation

A clear and comprehensive reference on the Wiener space is the book by Bogachev [8]

(see also [27]). We follow here closely the notation of [4]. Let X be a separable Banach space and let X

be its dual. We say that X is a Wiener space if it is endowed with a non-degenerate centered Gaussian probability measure γ. That amounts to say that γ is a probability measure for which x

♯γ is a centered Gaussian measure on R for every x

∈ X

. The non-degeneracy hypothesis means that γ is not concentrated on any proper subspace of X.

As a consequence of Fernique’s Theorem [8, Th. 2.8.5], for every x

∈ X

, the function R

x

(x) = h x

, x i is in L

2γ

(X) = L

2

(X, γ). Let H be the closure of R

X

in L

2γ

(X); the space H is usually called the reproducing kernel of γ. Let R, the operator from H to X, be the adjoint of R

that is, for ˆ h ∈ H ,

R ˆ h = Z

X

x ˆ h(x) dγ

where the integral is to be intended in the Bochner sense. It can be seen that R is a compact and injective operator. We will let Q = RR

. We denote by H the space R H . This space is called the Cameron-Martin space. It is a separable Hilbert space with the scalar product given by

[h

1

, h

2

]

H

= h ˆ h

1

, ˆ h

2

i

L2γ(X)

if h

i

= R ˆ h

i

. We will denote by | · |

H

the norm in H. The space H is a dense subspace of X, with compact embedding, and γ(H) = 0 if X is of infinite dimension.

For x

1

, .., x

m

∈ X

we denote by Π

x1,..,xm

the projection from X to R

m

given by Π

x1,..,xm

(x) = ( h x

1

, x i , .., h x

m

, x i ).

We will also denote it by Π

m

when specifying the points x

i

is unnecessary. Two elements x

1

and x

2

of X

will be called orthonormal if the corresponding h

i

= Qx

i

are orthonormal in H. We will fix in the following an orthonormal base of H given by h

i

= Qx

i

.

We also denote by H

m

= span(h

1

, .., h

m

) ≃ R

m

and X

m

= Ker(Π

m

) = H

mX

, so that X = R

m

⊕ X

m

. The map Π

m

induces the decomposition γ = γ

m

⊗ γ

m

, with γ

m

, γ

m

Gaussian measures on R

m

, X

m

respectively.

Proposition 2.1 ([8]). Let ˆ h

1

, .., ˆ h

m

be in H then the image measure of γ under the map Π

hˆ

1,..,ˆhm

(x) = (ˆ h

1

(x), .., ˆ h

m

(x))

is a Gaussian in R

m

. If the ˆ h

i

are orthonormal, then such measure is the standard

Gaussian measure on R

m

.

(5)

Given u ∈ L

2γ

(X), we will consider the canonical cylindrical approximation E

m

given by E

m

u(x) =

Z

Xm

u(Π

m

(x), y) dγ

m

(y).

Notice that E

m

u is a cylindrical functions depending only on the first m variables, and E

m

u converges to u in L

2γ

(X).

We will denote by FC

b1

(X) the space of cylindrical C

1

bounded functions that is the functions of the form v(Π

m

(x)) with v a C

1

bounded function from R

m

to R . We denote by FC

b1

(X, H) the space generated by all functions of the form Φh, with Φ ∈ FC

b1

(X) and h ∈ H.

We now give the definitions of gradients, Sobolev spaces functions of bounded variation.

Given u : X → R and h = R ˆ h ∈ H, we define

∂u

∂h (x) = lim

t→0

u(x + th) − u(x) t

whenever the limit exists, and

h

u = ∂u

∂h − ˆ h(u).

We define ∇

H

u : X → H, the gradient of u by

H

u =

+∞

X

i=1

∂u

∂h

i

h

i

and the divergence of Φ : X → H by div

γ

Φ =

+∞

X

i=1

hi

[Φ, h

i

]

H

.

The operator div

γ

is the adjoint of the gradient so that for every u ∈ FC

b1

(X) and every Φ ∈ FC

b1

(X, H), the following integration by parts holds:

Z

X

u div

γ

Φ dγ = − Z

X

[ ∇

H

u, Φ]

H

dγ. (1)

The ∇

H

operator is thus closable in L

2γ

(X) and we will denote by H

γ1

(X) its closure in L

2γ

(X). From this, formula (1) still holds for u ∈ H

γ1

(X) and Φ ∈ FC

b1

(X, H).

Following [21, 4], given u ∈ L

1γ

(X) we say that u ∈ BV

γ

(X) if Z

X

| D

γ

u |

H

= sup Z

X

u div

γ

Φ dγ; Φ ∈ FC

b1

(X, H), | Φ |

H

≤ 1 ∀ x ∈ X

< + ∞ .

(6)

We will also denote by | D

γ

u | (X) the total variation of u. If u = χ

E

is the characteristic function of a set E we will denote P

γ

(E) its total variation and say that E is of finite perimeter if P

γ

(E) is finite. As shown in [4] we have the following properties of BV

γ

(X) functions.

Theorem 2.2. Let u ∈ BV

γ

(X) then the following properties hold:

• D

γ

u is a countably additive measure on X with finite total variation and values in H (we will note the space of these measures by M (X, H)), such that for every Φ ∈ FC

b1

(X) we have:

Z

X

u ∂

hj

Φ dγ = − Z

X

Φdµ

j

∀ j ∈ N where µ

j

= [h

j

, D

γ

u]

H

.

• | D

γ

u | (X) = inf lim { R

X

|∇

H

u

i

|

H

dγ : u

j

∈ H

γ1

(X), u

j

→ u in L

1γ

(X) } .

Proposition 2.3. Let u = v(Π

m

) be a cylindrical function then u ∈ BV

γ

(X) if and only if v ∈ BV

γm

( R

m

). We then have

Z

X

| D

γ

u |

H

= Z

Rm

| D

γm

v | .

Proposition 2.4 (Coarea formula [2]). If u ∈ BV

γ

(X) then for every borel set B ⊂ X,

| D

γ

u | (B) = Z

R

P

γ

( { u > t } , B) dt. (2) In Proposition 3.12, we will need the following extension of Proposition 2.4.

Lemma 2.5. For every function u ∈ BV

γ

(X) and every non-negative Borel function g, Z

X

g(x) d | D

γ

u | (x) = Z

R

Z

X

g(x) d | D

γ

χ

Et

| (x)

dt (3)

where E

t

:= { u > t } .

Proof. The proof of this lemma mimic the standard proof in the Euclidean case [14, Th.2.2]. By [20, Ch.1,Th.7] we can write g as

g =

+∞

X

i=1

1

i χ

Ai

(7)

where the A

i

⊂ X are Borel sets. Using the coarea formula (2), we then get Z

X

g(x)d | D

γ

u | (x) = X

+∞

i=1

1

i | D

γ

u | (A

i

)

=

+∞

X

i=1

1 i

Z

R

| D

γ

χ

Et

| (A

i

) dt

= Z

R

Z

X

X

+∞

i=1

1

i χ

Ai

d | D

γ

χ

Et

| (x)

! dt

= Z

R

Z

X

g(x) d | D

γ

χ

Et

| (x) dt.

In [4] it is also shown that sets with finite Gaussian perimeter can be approximated by smooth cylindrical sets.

Proposition 2.6. Let E ⊂ X be a set of finite Gaussian perimeter then there exists smooth sets E

m

⊂ R

m

such that Π

m1

(E

m

) converges in L

1γ

(X) to E and P

γ

m1

(E

m

)) = P

γm

(E

m

) converges to P

γ

(E) when m tends to infinity.

Note that, for half-spaces, the perimeter can be exactly computed [4, Cor. 3.11].

Proposition 2.7. Let h = R h ˆ ∈ H and c ∈ R then the half-space E = { x ∈ X : ˆ h(x) ≤ c }

has perimeter

P

γ

(E) = 1

√ 2π e

c2 2|h|2

H

.

3 The Ehrhard symmetrization

The Ehrhard symmetrization has been introduced by Ehrhard in [18] for studying the isoperimetric inequality in a Gaussian setting. We recall the definition and the main properties of such symmetrization.

Definition 3.1. We define the functions Φ and α by Φ(x) = 1

√ 2π Z

x

−∞

e

t

2

2

dt and α(x) = Φ

−1

(x).

we then let U (x) = Φ

◦ α(x) =

1

e

α2(x)2

.

(8)

Notice that Φ(t) is the volume of the half-space { ˆ h(x) < t } and that U (x) is the perimeter of a half-space of volume x.

Lemma 3.2. Let ˆ h

1

, ˆ h

2

∈ H , with | h

1

|

H

= | h

2

|

H

= 1, and suppose that there exist C

1

, C

2

∈ R such that

{ ˆ h

1

< C

1

} ⊂ { h ˆ

2

< C

2

} . Then h ˆ

1

= ˆ h

2

.

Proof. Assume by contradiction ˆ h

1

6 = ˆ h

2

, and let η > 0 be such that | ˆ h

1

− ˆ h

2

|

L2γ(X)

≥ η.

We shall bound from below by a positive constant the quantity γ n

h ˆ

1

(x) < C

1

o

∩ n

ˆ h

2

(x) ≥ C

2

o thus contradicting the inclusion

n ˆ h

1

< C

1

o

⊂ n

ˆ h

2

< C

2

o .

Letting h be a unitary vector in H orthogonal to h

1

, we can write h

2

= λh

1

+ βh

with λ

2

+ β

2

= 1. Up to exchanging h with − h, we can also assume that β ≥ 0. We then have | h

1

− h

2

|

H

= 2(1 − λ) and thus − 1 ≤ λ ≤ 1 −

η2

.

Let us first suppose that − 1 ≤ λ ≤ −

12

, then

h ˆ

1

(x) < min

C

1

, − C

2

λ

∩ n

h(x) ˆ ≥ 0 o

⊂ n

ˆ h

1

(x) < C

1

o

∩ n

ˆ h

2

(x) ≥ C

2

o . As ˆ h

1

and ˆ h are orthogonal we have Π

ˆh

1,ˆh

♯γ = γ

2

and thus γ

ˆ h

1

(x) < min(C

1

, − C

2

λ )

∩ n

ˆ h(x) ≥ 0 o

= 1

2 Φ(min(C

1

, − C

2

/λ))

≥ 1

2 Φ(min(C

1

, 2C

2

)).

Hence, for − 1 ≤ λ ≤ −

12

, γ n

ˆ h

1

(x) < C

1

o

∩ n

h ˆ

2

(x) ≥ C

2

o

≥ 1

2 Φ(min(C

1

, 2C

2

)).

(9)

If now −

12

≤ λ ≤ 1 −

η2

, we can assume that η is such that 1 −

η2

12

. Let us start by computing the Fourier transform of Π

hˆ

1,ˆh2

♯γ. Denoting by ˜ µ the Fourier transform of a measure µ (see [8, Sec. 1.2]) and letting Π := Π

ˆh

1,ˆh2

, for every (z

1

, z

2

) ∈ R

2

we have Π♯γ(z g

1

, z

2

) =

Z

R2

e

iz·x

dΠ♯γ(x)

= Z

X

e

iz·Π(x)

dγ(x)

= Z

X

e

i[z1ˆh1(x)+z2hˆ2(x)]

dγ(x)

= Z

X

e

i[(z1+z2λ)ˆh1(x)+z2βˆh(x)]

dγ(x)

= Z

R2

e

i[(z1+z2λ)x1+z2βx2]

2

(x

1

, x

2

)

= γ e

2

(z

1

+ λz

2

, βz

2

)

= e

12[(z1+λz2)22z22]

= e

12[z12+z22+2λz1z2]

. Thus, if we set K :=

1 λ λ 1

, we have Π♯γ(z) = g e

12ztKz

. It follows that Π♯γ is a centered Gaussian measure with density

1

2π√

detK

e

12ztK−1z

and thus Π♯γ(z

1

, z

2

) =

√ 1 − λ

2

2π e

12[z12+z22−2λz1z2]

dz.

We now compute γ n

ˆ h

1

(x) < C

1

o

∩ n

ˆ h

2

(x) ≥ C

2

o

= Z

X

χ

{ˆh

1(x)<C1}

(x)χ

{hˆ

2(x)≥C2}

(x) dγ(x)

= Z

R2

χ

{z1<C1}

(z)χ

{z2C2}

(z) dΠ♯γ(z)

= Z

C1

−∞

Z

+∞

C2

√ 1 − λ

2

2π e

12[z21+z222λz1z2]

dz

1

dz

2

≥ 1 2π

r 3 4

Z

C1

−∞

Z

+∞

C2

e

12z12

e

12z22

e

λz1z2

dz

1

dz

2

. Finally, when λz

1

z

2

≥ 0, we can bound e

λz1z2

from below by 1, and when λz

1

z

2

≤ 0 we can bound it form below by e

12|z1z2|

so that we can always bound from below

γ

{ ˆ h

1

(x) < C

1

} ∩ { ˆ h

2

(x) ≥ C

2

}

(10)

by a positive constant.

We now define the Ehrhard symmetrization.

Definition 3.3. Let E ⊂ X and let m ∈ N . The Ehrhard symmetral of E along the first m variables is defined as (see Figure 1):

E

:=

 

(x, x

m

, x

m

) ∈ R

m−1

× R × X

m

: x

m

< α( E

m1

χ

E

(x)) if m > 1 { x ∈ X : h x

1

, x i < α(γ (E)) } if m = 1.

x

m

E

m1

χ

E

(x) E

E

s

E

m1

χ

E

(x) α( E

m−1

χ

E

)

x

Figure 1: The Ehrhard symmetrization.

The interest of this symmetrization is that it decreases the Gaussian perimeter, while keeping the volume fixed.

Proposition 3.4. Let E be a set of finite perimeter and E

be an Ehrhard symmetral of E, then

γ(E

) = γ(E), (4)

E

m1

χ

E

= E

m1

χ

E

and

P

γ

(E

) ≤ P

γ

(E). (5)

(11)

In particular, we have the isoperimetric inequality P

γ

(E) ≥ U (γ (E)), with equality if and only if E is a half-space.

For the proof we refer to [7, 10], and to [4] for the extension to infinite dimensions.

We can also prove a stronger result which is a kind of Bernstein Theorem in this setting.

Proposition 3.5. The half-spaces are the only local minimizers of the Gaussian perimeter with volume constraint.

Proof. Let E ⊂ X be a local minimizer of the (Gaussian) perimeter and let v = γ (E).

This means that, for every R > 0 and every set F of finite perimeter, with γ (F) = v and E∆F ⊂ B

R

(where B

R

denotes the ball of radius R centered at 0), we have

P

γ

(E) ≤ P

γ

(F ).

If E is not an half space then, by Proposition 3.4, there exists η > 0 such that P

γ

(E) ≥ U (v) + η.

Let α

R

be such that

γ (E ∩ B

R

) = γ ( {h x

1

, x i < α

R

} ∩ B

R

) .

We have that α

R

tends to α(v) when R goes to infinity and P

γ

( {h x

1

, x i < α

R

} ) tends to P

γ

( {h x

1

, x i < α(v) } ). Letting

F

R

= ( {h x

1

, x i < α

R

} ∩ B

R

) ∪ (E ∩ B

Rc

) we get

U (v) + η ≤ P

γ

(E) ≤ P

γ

(F

R

) ≤ P

γ

( {h x

1

, x i < α

R

} ∩ B

R

) + P

γ

(E ∩ B

Rc

)

≤ P

γ

( {h x

1

, x i < α

R

} ) + P

γ

(B

R

) + P

γ

(E ∩ B

Rc

)

≤ P

γ

( {h x

1

, x i < α(v) } ) + ε(R)

= U (v) + ε(R), where we used various time the inequality (see [23])

P

γ

(E ∪ F ) + P

γ

(E ∩ F) ≤ P

γ

(E) + P

γ

(F )

and where ε(R) is a function which goes to zero when R goes to infinity. We thus found

a contradiction.

(12)

Remark 3.6. In the Euclidean setting, half-spaces are the only local minimizers of the perimeter only in dimension lower than 8 (see [23]). Notice also that if we drop the volume constraint, half spaces are no longer local minimizers for the Gaussian perimeter, since there are no nonempty local minimizers.

In the sequel we will also need another transformation which from a finite dimensional function gives an Ehrhard symmetric set whose sections have volume prescribed by the original function. More precisely:

Definition 3.7. Given a measurable function v : R

m

→ [0, 1], we define its Ehrhard set ES

m

(v) ⊂ X by

ES

m

(v) := n

(x, x

m+1

, x

m+1

) ∈ R

m

× R × X

m+1

: x

m+1

< α(v(x)) o .

Given a measurable cylindrical function u : X → [0, 1] depending only on the first m variables, that is, u = v ◦ Π

m

for some v : R

m

→ [0, 1], we set

ES

m

(u) := ES

m

(v).

The link between Ehrhard sets and Ehrhard symmetrization is the following:

Proposition 3.8. Let E be a set of finite perimeter and E

be its Ehrhard symmetrization with respect to the first (m + 1) variables, then

E

= ES

m

( E

m

E

)).

In the next proposition we compute the perimeter of Ehrhard sets. It slightly extends a result in [15].

Proposition 3.9. Let u ∈ BV

γm

( R

m

) with 0 ≤ u ≤ 1, then P

γ

(ES

m

(u)) =

Z

Rm

q

U (u)

2

+ | D

γm

u |

2

m

where Z

Rm

q

U (u)

2

+ | D

γm

u |

2

m

= Z

Rm

p U (u)

2

+ |∇ u |

2

m

+ | D

γs

u | (X)

and D

γ

u = ∇ u γ + D

γs

u is the Radon-Nikodym decomposition of D

γ

u.

Proof. By [15, Th. 4.3] the result holds for u ∈ H

γ1m

( R

m

). We will show by approximation that the same holds for u ∈ BV

γm

( R

m

).

Let E = ES

m

(u), then we can find sets E

n

such that γ (E

n

∆E) → 0 and P

γ

(E

n

) → P

γ

(E)

as n → + ∞ , and all the E

n

have smooth boundary and are Ehrhard symmetric. Thus,

(13)

for every n ∈ N , there exists a smooth function u

n

such that 0 ≤ u

n

≤ 1, E

n

= ES

m

(u

n

), u

n

→ u in L

1γm

( R

m

), and

P

γ

(E

n

) = Z

Rm

q

U (u

n

)

2

+ | D

γm

u

n

|

2

m

. Since, by Proposition 4.4, the functional R

Rm

p U (u)

2

+ | D

γm

u |

2

m

is lower semicontin- uous in L

1γm

( R

m

), we get

P

γ

(E) = lim

n→∞

P

γ

(E

n

)

= lim

n→∞

Z

Rm

q

U (u

n

)

2

+ | D

γm

u

n

|

2

m

≥ Z

Rm

q

U (u)

2

+ | D

γm

u |

2

m

.

The other inequality follows as in [15]. Let E e = Π

m+1

(E) ⊂ R

m+1

and observe that γ

m+1

( E) = e γ (E) and P

γm+1

( E) = e P

γ

(E). By Vol’pert Theorem [2, Th. 3.108] there exists a set B ⊂ R

m

such that for every x ∈ B , ν

m+1Ee

(x, α(u

E

(x))) exists and is not equal to zero, where ν

m+1Ee

denotes the last coordinate of the unit external normal to ∂

E. By e [15, Lemma 4.4], γ

m

-almost every x ∈ B is a point of approximate differentiability for u.

By Lemma 4.5 and 4.6 of [15] we then have

P

γm+1

( E) = e P

γm+1

( E, B e × R ) + P

γm+1

( E, B e

c

× R )

≤ Z

B

p U (u)

2

+ |∇ u |

2

m

+ Z

Bc

| D

γm

u | + Z

Bc

U (u) dγ

m

. As γ

m

(B

c

) = 0, we find that

Z

B

p U (u)

2

+ |∇ u |

2

m

+ Z

Bc

| D

γm

u | = Z

Rm

p U

2

(u) + |∇ u |

2

m

+ | D

γsm

u | ( R

m

)

and thus P

γ

(E) = P

γm+1

( E) e ≤ R

Rm

p U (u)

2

+ | D

γm

u |

2

m

.

The last transformation that we consider is the analog of the Schwarz symmetrization in the Gaussian setting, and was first introduced by Ehrhard in [19].

Definition 3.10. Let u ∈ X → R be a measurable function and let m ∈ N be fixed. We define the m-dimensional Ehrhard symmetrization u

of u as follows:

• for all t ∈ R we let E

t

be the Ehrhard symmetrization of E

t

:= { u > t } with respect

to the first m variables;

(14)

• we let u

(x) := inf { t : x ∈ E

t

} .

As (4) implies γ ( { u

> t } ) = γ ( { u > t } ) for all t ∈ R , from the Layer Cake formula it follows that, if u ∈ L

2γ

(X), then u

∈ L

2γ

(X) and

Z

X

| u

|

2

dγ = Z

X

| u |

2

dγ . (6)

Indeed, we have Z

X

| u |

2

dγ = 2 Z

+∞

0

t γ( { u > t } ) dt − 2 Z

0

−∞

t γ( { u < t } ) dt

= 2 Z

+∞

0

t γ( { u

> t } ) dt − 2 Z

0

−∞

t γ( { u

< t } ) dt

= Z

X

| u

|

2

dγ.

Lemma 3.11. Let u, v : X → [0, + ∞ ) belonging to L

2γ

(X), then

k u

− v

k

L2γ(X)

≤ k u − v k

L2γ(X)

. (7) Proof. The proof is a straightforward adaptation of the analogous proof for the Schwarz symmetrization [25, Th. 3.4].

Recalling (6) with p = 2, we have only to show that Z

X

uvdγ ≤ Z

X

u

v

dγ. (8)

Again by the Layer Cake formula we have Z

X

uvdγ = Z

+∞

0

Z

+∞

0

Z

X

χ

{u>t}

(x)χ

{v>s}

(x)dγ(x) dt ds.

Thus (8) would follow from the same inequality for sets, that is,

γ (A ∩ B) ≤ γ (A

∩ B

) . (9) Let x

m

∈ R

m

and assume that

Z

Xm

χ

A

(x

m

+ y)dγ

m

(y) ≥ Z

Xm

χ

B

(x

m

+ y)dγ

m

(y) then by definition of the Ehrhard symmetrization we have

B

∩ (x

m

+ X

m

) ⊂ A

∩ (x

m

+ X

m

)

(15)

and therefore Z

Xm

χ

A

(x

m

+ y)χ

B

(x

m

+ y)dγ

m

(y) = Z

Xm

χ

A

(x

m

+ y)dγ

m

(y)

= Z

Xm

χ

A

(x

m

+ y)dγ

m

(y)

≥ Z

Xm

χ

A

(x

m

+ y)χ

B

(x

m

+ y)dγ

m

(y) This inequality also holds if R

Xm

χ

B

(x

m

+ y)dγ

m

(y) ≥ R

Xm

χ

A

(x

m

+ y)dγ

m

(y) so that finally

γ (A

∩ B

) = Z

Xm

Z

Xm

χ

A

(x + y)χ

B

(x + y)dγ

m

(y)dγ

m

(x)

≥ Z

Xm

Z

Xm

χ

A

(x + y)χ

B

(x + y)dγ

m

(y)dγ

m

(x)

= γ (A ∩ B) which gives (9).

As for the Schwarz symmetrization, a P´ olya-Szeg¨ o principle holds for the Ehrhard sym- metrization.

Proposition 3.12. Let u ∈ H

γ1

(X), let m ∈ N and let u

be the m-dimensional Ehrhard symmetrization of u. Then u

∈ H

γ1

and

Z

X

|∇

H

u

|

2H

dγ ≤ Z

X

|∇

H

u |

2H

dγ. (10)

Moreover, if m = 1 and equality holds in (10), then u = ˜ u

ˆ h(x)

for some ˆ h ∈ H , and ˆ h can be chosen to be a unitary vector.

Proof. In [19, Th. 3.1], inequality (10) is proven for Lipschitz functions, in finite dimen- sions. We extend it by approximation to Sobolev functions.

We can assume u ≥ 0, since we have (u

±

)

= (u

)

±

, where u

±

, (u

)

±

denote the positive and negative part of u and u

, respectively.

Let u

n

∈ FC

c1

(X) be positive functions converging to u in H

γ1

(X), then by (7), u

n

converges to u

in L

2γ

(X) and thus by the lower semicontinuity of the H

γ1

(X) norm we

have Z

X

|∇

H

u

|

2H

≤ lim

n→∞

Z

X

|∇

H

u

n

|

2H

≤ lim

n→∞

Z

X

|∇

H

u

n

|

2H

= Z

X

|∇

H

u |

2H

.

(16)

We now turn to the equality case for one-dimensional symmetrizations. For this we closely follow [10] and give an alternative proof of (10), based on ideas of Brothers and Ziemer [9] for the Schwarz symmetrization.

Let u ∈ H

γ1

(X) and µ(t) = γ( { u > t } ) = γ ( { u

> t } ). By the coarea formula (3), for all t ∈ R we have

µ(t) = γ( { u > t } ∩ {∇

H

u = 0 } ) + Z

+∞

t

Z

{∇Hu6=0}

1

|∇

H

u |

H

d | D

γ

χ

Eτ

|

! dτ.

Hence

− µ

(t) ≥ Z

{∇Hu6=0}

1

|∇

H

u |

H

d | D

γ

χ

Et

| for a.e. t ∈ R . (11) Since u

is a function depending only on one variable, arguing as in [14] we get

d

dt γ ( { u

> t } ∩ {∇

H

u

= 0 } ) = 0 for a.e. t ∈ R .

As u

is monotone we have that |∇

H

u

|

H

is constant on { u

= t }∩{∇

H

u

6 = 0 } . Observe also that, being u

one-dimensional, { u

= t } has a well defined meaning. We thus find:

− µ

(t) = P

γ

( { u

> t } )

|∇

H

u

|

{u=t}

for a.e. t ∈ R , which implies, recalling (11),

P

γ

( { u

> t } )

|∇

H

u

|

{u=t}

≥ Z

{∇Hu6=0}

1

|∇

H

u |

H

d | D

γ

χ

Et

| for a.e. t ∈ R . (12) Let us note that as in [10, Lem. 4.2], using (3) with g = χ

{∇Hu=0}

we find

Z

X

χ

{∇Hu=0}

|∇

H

u |

H

dγ = 0 = Z

R

Z

X

χ

{∇Hu=0}

d | D

γ

χ

Et

| (x) dt and thus for almost every t ∈ R ,

Z

X

χ

{∇Hu=0}

d | D

γ

χ

Et

| (x) = 0.

This shows that for almost every t ∈ R , ∇

H

u(x) 6 = 0 for | D

γ

χ

Et

| -almost every x ∈ X and thus

Z

{∇Hu6=0}

1

|∇

H

u |

H

d | D

γ

χ

Et

| (x) = Z

X

1

|∇

H

u |

H

d | D

γ

χ

Et

| (x) for a.e. t ∈ R . (13)

(17)

By (3), (5),(12) and (13), we eventually get Z

X

|∇

H

u

|

2

dγ = Z

R

|∇

H

u

|

{u=t}

P

γ

( { u

> t } )dt

= Z

R

P

γ

( { u

> t } )

2

P

γ({u>t})

|∇Hu|{u=t}

dt

≤ Z

R

P

γ

( { u > t } )

2

R

X 1

|∇Hu|H

d | D

γ

χ

Et

| (x) dt

≤ Z

R

Z

X

|∇

H

u |

H

d | D

γ

χ

Et

| (x) dt

= Z

X

|∇

H

u |

2H

dγ .

As a consequence, if equality holds in (10), then equality holds in the Gaussian isoperi- metric inequality, that is,

P

γ

(u > t) = P

γ

(u

> t) for a.e. t ∈ R .

This implies that almost every level-set of u is a half-space, i.e. for almost every t ∈ R there exists ˆ h

t

∈ H such that { u > t } = { ˆ h

t

< α(µ(t)) } , and without loss of generality we can assume that | h

t

|

H

= 1. Such half-spaces being nested, by Lemma 3.2 we have that ˆ h

t

does not depend on t and thus u(x) = v(ˆ h(x)).

Remark 3.13. We notice that the fact that equality in (10) implies that u is one- dimensional is a specific feature of the Gaussian setting, and the analogous statement does not hold for the Schwarz symmetrization in the Euclidean case [9]. Indeed, this property is a consequence of the fact that Gaussian measures, differently from the Lebesgue measure, are not invariant under translations.

4 Relaxation of perimeter

In this section we compute the relaxation of the perimeter functional F (u) :=

( P

γ

(E) if u = χ

E

+ ∞ otherwise

(18)

with respect to the weak L

2γ

(X)-topology. The fact that F is not lower semicontinuous can be easily checked by taking the sequence E

n

= {h x

n

, x i < 0 } . Indeed, the characteristic functions of these sets weakly converge to the constant function 1/2, which is not a characteristic function, while the perimeter of E

n

is constantly equal to 1/ √

2π.

We will show that the relaxation of F is equal to F(u) :=

 

 Z

X

q

U

2

(u) + | D

γ

u |

2

dγ if 0 ≤ u ≤ 1 γ − a.e.

+ ∞ otherwise

where Z

X

q

U

2

(u) + | D

γ

u |

2

dγ = Z

X

q

U

2

(u) + |∇

H

u |

2H

dγ + | D

sγ

u | (X)

with D

γ

u = ∇

H

udγ + D

γs

u. Observe that the functional F already appears in the seminal work of Bakry and Ledoux [6] and in the earlier work of Bobkov [7] in the context of log- Sobolev inequalities. This functional has been also studied in [10]. See also [4, Remark 4.3] where it appears in a setting closer to ours.

Let us first recall the definition of the lower semicontinous envelope of a function (see [16] for more details).

Definition 4.1. Let X be a topological vector space. For every function F : X → R , its lower semicontinuous envelope (or relaxed function) is the greatest lower semicontinuous function that lies below F .

When X is a metric space, the following caracterization holds.

Proposition 4.2. Let X be a metric space. For every function F : X → R , and every x ∈ X, the relaxed function F is given by

F (x) = inf

lim

n→∞

F (x

n

) : x

n

→ x

.

We now show a representation formula for F which is reminiscent of the definition of the total variation and of the nonparametric area functional (see [23]). We start with a preliminary result.

Lemma 4.3. Let g ∈ L

(X) with g ≥ 0, let µ ∈ M (X, H), and define f(g, µ) := ˜

q

g

2

+ | h |

2H

dγ + | µ

s

| , where µ = h γ + µ

s

. There holds

f ˜ (g, µ)(X) = sup

Φ∈L1 µ(X,H)

ξ∈L1µ(X)

Z

X

[Φ, dµ]

H

+ Z

X

g ξ dγ : | Φ |

2H

+ | ξ |

2

≤ 1 a.e. in X

. (14)

(19)

Proof. The proof is adapted from [17].

Notice first that, for (λ, p) ∈ R × H, the function f (λ, p) := q

λ

2

+ | p |

2H

defines a norm on the product space R × H. Moreover, if we let f

λ

(p) := q

λ

2

+ | p |

2H

, then the convex conjugate of f

λ

is f

λ

(Φ) = − λ q

1 − | Φ |

2H

. We divide the proof into three steps.

Step 1. Let

M (g, µ) = sup

Φ∈L1µ(X,H)

Z

X

[Φ, dµ]

H

+ Z

X

g q

1 − | Φ |

2H

dγ : | Φ |

H

≤ 1 a.e. in X

. We will show that

M (g, hγ ) = Z

X

f (g, h)dγ. (15)

By definition of convex conjugate, it is readily checked that M (g, hγ ) ≤ R

X

f (g, h)dγ. We thus turn to the other inequality. By definition of the Bochner integral, for every δ > 0, there exists h

i

∈ H and A

i

⊂ X with A

i

disjoints Borel sets and i ∈ [1, m] such that if we set

θ = X

m

i=1

χ

Ai

h

i

then | θ − h |

L1γ

≤ δ. Analogously there exists η

i

∈ X such that setting

˜ g =

X

m

i=1

χ

Ai

η

i

we have | ˜ g − g |

L1γ

≤ δ. By the observation at the beginning of the proof and the triangle inequality we get

| f (˜ g, θ) − f (g, h) | ≤ f (˜ g − g, θ − h) | ≤ | g ˜ − g | + | θ − h |

H

.

For every i, by definition of convex conjugate, there exists ξ

i

∈ H with | ξ

i

|

H

≤ 1 such that

f (η

i

, h

i

) ≤ [ξ

i

, h

i

]

H

+ η

i

q

1 − | ξ

i

|

2H

+ δ.

(20)

From this, setting Φ = P

m

i=1

χ

Ai

ξ

i

we have Z

X

f(g, h)dγ ≤ Z

X

f (˜ g, θ)dγ + 2δ

= X

m

i=1

Z

Ai

f (η

i

, h

i

)dγ + 2δ

≤ X

m

i=1

Z

Ai

i

, h

i

]

H

+ η

i

q

1 − | ξ

i

|

2H

dγ + 3δ

= Z

X

[Φ, h]

H

+ ˜ g q

1 − | Φ |

2H

dγ + 3δ.

Since ˜ g q

1 − | Φ |

2H

− g q

1 − | Φ |

2H

≤ | g ˜ − g | we find Z

X

f(g, h)dγ ≤ Z

X

Φ · h − g q

1 − | Φ |

2H

dγ + 4δ

≤ M(g, hγ) + 4δ.

Since δ is arbitrary we have M (g, hγ) = R

X

f (g, h)dγ.

Step 2. The proof proceeds exactly as in [17] and we only sketch it. Recalling (15), it remains to show that

M (g, hγ + µ

s

) = M (g, hγ) + | µ

s

| (X).

One inequality is easily obtained, since M (g, hγ + µ

s

) = sup

Φ

Z

X

[Φ, h]

H

dγ + Z

X

Φ · dµ

s

+ Z

X

g(x) q

1 − | Φ |

2H

sup

Φ

Z

X

[Φ, h]

H

dγ + Z

X

g(x) q

1 − | Φ |

2H

+ Z

X

| dµ

s

|

= M (g, hγ) + | µ

s

| (X).

For the opposite inequality, let δ > 0 be fixed then there exists Φ

1

and Φ

2

such that M (g, hγ) ≤

Z

X

1

, h]

H

dγ + Z

X

g(x) q

1 − | Φ

1

|

2H

dγ + δ

| µ

s

| (X) ≤ Z

X

2

, dµ

s

]

H

+ δ.

(21)

Taking Φ equal to Φ

2

on a sufficiently small neighborhood of the support of µ

s

and equal to Φ

1

outside this neighborhood, we get

M(g, hγ) + | µ

s

| (X) ≤ Z

X

[Φ, h]

H

dγ + Z

X

g(x) q

1 − | Φ |

2H

dγ + Z

X

[Φ, dµ

s

]

H

+ Cδ

≤ M(g, hγ + µ

s

) + Cδ which gives the opposite inequality.

Step 3. In order to conclude the proof, it is enough to notice that for every Φ ∈ L

1µ

(X, H), with | Φ |

H

≤ 1, we have

sup

ξ∈L1µ(X)

Z

X

[Φ, dµ]

H

+ Z

X

g ξ dγ : | Φ |

2H

+ | ξ |

2

≤ 1 a.e. in X

= Z

X

[Φ, dµ]

H

+ Z

X

g q

1 − | Φ |

2H

dγ.

Proposition 4.4. Let u ∈ BV

γ

(X) then F (u) = sup

Φ∈F C1 b(X,H)

ξ∈FC1b(X)

Z

X

(u div

γ

Φ + U (u)ξ) dγ : | Φ(x) |

2H

+ | ξ(x) |

2

≤ 1 ∀ x ∈ X

. (16)

Proof. We apply Lemma 4.3 with µ = Du and g = U (u). Since µ is tight [4], the space FC

b1

(X, H) is dense in L

1µ

(X, H) so that we can restrict the supremum in (16) to smooth cylindrical functions Φ, ξ.

Remark 4.5. Since U is concave, the duality formula (16) is not sufficient to prove that F is lower semicontinuous for the weak L

2γ

(X)-topology. It shows however the lower- semicontinuity of F in the strong L

2γ

(X)-topology.

We now prove that F is the lower semicontinuous envelope of F . Theorem 4.6. F is the relaxation of F in the weak L

2γ

(X)-topology.

Proof. Let us first notice that F takes finite values only on functions of the closed unit ball of L

2γ

(X) which is metrizable for the weak convergence. Therefore the relaxation and the sequential relaxation in the weak topology of L

2γ

(X) coincide.

Let χ

En

be a sequence of sets weakly converging in L

2γ

(X) to u ∈ BV

γ

(X), with uniformly bounded perimeter. We shall show that

lim

n→∞

P

γ

(E

n

) ≥ F(u).

(22)

Notice that, by weak convergence, we necessarily have 0 ≤ u ≤ 1 a.e. on X.

For all n ≥ 1 and k ≥ 2, we let E

nk

be the Ehrhard symmetral of E

n

with respect to the first k variables. Recalling the notation of Section 3, we have

P

γ

(E

nk+1

) ≤ P

γ

(E

n

) and E

nk+1

= ES

k

( E

k

χ

En

) . As

Z

X

| D

γ

E

k

En

) |

H

≤ P

γ

(E

n

) and E

k

En

) depends only on the first k variables, by the compact embedding of BV

γk

( R

k

) into L

1γk

( R

k

) we can extract a subsequence from E

k

En

) which converges strongly to u

k

:= E

k

(u). From this we get that E

nk+1

= ES

k

( E

k

χ

En

) tends strongly to E

k+1

:= ES

k

(u

k

). By the lower semicontinuity of the perimeter we then have

lim

n→∞

P

γ

(E

n

) ≥ lim

n→∞

P

γ

(E

nk+1

) ≥ P

γ

(E

k+1

).

For every ϕ ∈ FC

b1

(X), with ϕ depending only of the j ≤ k first variables, there holds Z

X

χ

Ek+1

(x)ϕ(x)dγ(x) = Z

X

u

k

(x)ϕ(x)dγ(x) = Z

X

u(x)ϕ(x)dγ(x),

which implies that the sequence χ

Ek+1

tends weakly to u. In order to conclude the proof it remains to show that

k

lim

→∞

P

γ

(E

k+1

) = F(u).

Notice that, by Proposition 3.9, there holds

P

γ

(E

k+1

) = F (u

k

).

For every Φ ∈ FC

b1

(X, H) and ξ ∈ FC

b1

(X), depending on the first k variables and such that the range of Φ is included in H

k

, by Proposition 4.4, we have

Z

X

u

k

div

γ

Φ + U (u

k

)ξ dγ =

Z

X

(u div

γ

Φ + U (u)ξ) dγ ≤ F (u).

Taking the supremum in Φ, ξ and recalling (16), we then get F (u

k

) ≤ F (u) for all k.

Repeating the same argument with u

k+1

instead of u, we obtain that F (u

k

) is nonde- creasing in k. Therefore there exists ℓ ≥ 0 such that

k

lim

→∞

F(u

k

) = lim

k→∞

P

γ

(E

k+1

) = ℓ ≤ F (u).

(23)

Assume by contradiction that ℓ < F (u). Then there exists δ > 0 such that F (u

k

) ≤ F (u) − δ for all k, hence there exist N ∈ N , Φ ∈ FC

b1

(X, H) and ξ ∈ FC

b1

(X), depending only on the first N variables, such that

Z

X

u

k

div

γ

Φ + U (u

k

dγ ≤ F (u

k

) ≤ F (u) − δ ≤ Z

X

(u div

γ

Φ + U (u)ξ) dγ − δ 2 , but for k > N we have

Z

X

u

k

div

γ

Φ + U (u

k

)ξ dγ =

Z

X

(u div

γ

Φ + U (u)ξ) dγ which leads to a contradiction.

Remark 4.7. Theorem 4.6 provides an example of a nonconvex functional, namely F , which is lower semicontinuous for the weak L

2γ

(X)-topology. We also know that semicon- tinuity does not holds for general functional of the form

J (u) = Z

X

f (u, D

γ

u)dγ since if we take for instance f (u, p) := p

g

2

(u) + | p |

2

with g such that g(1/2) > U (1/2) and g(0) = g(1) = 0, then, letting u

n

:= {h x

n

, x i < 0 } , we have u

n

⇀ u = 1/2 weakly in L

2γ

(X), so that

J(u) = g 1

2

> U 1

2

= 1

√ 2π = lim

n→∞

J (u

n

).

One could wonder what are the right hypotheses for a functional of this form to be lower semicontinuous with respect to the weak topology.

5 Γ-limit for the Modica-Mortola functional

Let us briefly recall the definition of Γ-convergence. We refer to [16] for a comprehensive treatment of the subject.

Definition 5.1. Let X be a topological space, and let F

n

: X → R be a sequence of functions. The Γ-lower limit and the Γ-upper limit of the sequence F

n

is defined as

(Γ − lim

n→∞

F

n

)(x) = sup

U∈N(x)

lim

n→∞

y

inf

∈U

F

n

(y) (Γ − lim

n→∞

F

n

)(x) = sup

U∈N(x)

n

lim

→∞

inf

y∈U

F

n

(y)

where N (x) denotes the set of all open neighbourhoods of x in X. When the Γ-lower limit

and the Γ-upper limit coincide, we say that the sequence F

n

Γ-converges.

(24)

As for the relaxation, if X is a metric space we have a sequential caracterization of the Γ-convergence.

Theorem 5.2. Let X be a metric space. A sequence of functions F

n

Γ-converges to F : X → R if and only if the following two conditions hold:

• for every sequence x

n

converging to x, it holds lim

n→∞

F

n

(x

n

) ≥ F (x)

• for every x ∈ X there exists a sequence x

n

converging to x with lim

n→∞

F

n

(x

n

) ≤ F (x).

Let now W ∈ C

1

( R ) be a double-well potential with minima in { 0, 1 } , that is, W (t) ≥ 0 for all t ∈ R , and W (t) = 0 iff t ∈ { 0, 1 } . We also assume W (t) ≥ C(t

2

− 1) for some C > 0 and t ∈ R . A typical example of such potential is W (t) = t

2

(t − 1)

2

.

For any ε > 0 we define the functionals F

ε

: L

2γ

(X) → [0, + ∞ ] as

F

ε

(u) :=

 

 

 

 Z

X

ε

2 |∇

H

u |

2H

+ W (u) ε

dγ if u ∈ H

γ1

(X)

+ ∞ if u ∈ L

2γ

(X) \ H

γ1

(X) . We are ready to prove our main Γ-convergence result.

Theorem 5.3. When ε tends to zero the functionals F

ε

Γ-converge, in the weak topology of L

2γ

(X), to the functional c

W

F , where c

W

= R

1

0

p 2W (t) dt.

Proof. Notice first that the Γ-limit does not change if we restrict the domain of F

ε

to the functions u ∈ H

γ1

(X) such that 0 ≤ u ≤ 1. This follows from the following two facts:

- for all u ∈ H

γ1

(X), letting ˜ u = min(max(u, 0), 1), we have F

ε

(˜ u) ≤ F

ε

(u);

- F

ε

(u) ≥ R

X W(u)

ε

dγ for all u ∈ H

γ1

(X), which implies that the Γ-limit is concentrated on the functions u ∈ L

2γ

(X) such that u(x) ∈ { 0, 1 } for a.e. x ∈ X.

Since the restricted domain is contained in the unit ball of L

2γ

(X), which is metrizable for the weak L

2γ

(X)-topology, by Theorem 5.2 the Γ-limit and the sequential Γ-limit of F

ε

coincide.

We now compute the Γ-liminf of F

ε

.

Let u

ε

∈ H

γ1

(X) be such that 0 ≤ u

ε

≤ 1 and F

ε

(u

ε

) ≤ C for some C > 0. Then R

X

W (u

ε

)dγ ≤ Cε, which gives a uniform bound on k u

ε

k

L2γ(X)

recalling that W (u) ≥

C(u

2

− 1). As a consequence, there exists a weakly converging subsequence, still denoted

(25)

by u

ε

. Letting u be its weak limit, from 0 ≤ u

ε

≤ 1 we get 0 ≤ u ≤ 1. Using the coarea formula (2), we obtain the estimate

F

ε

(u

ε

) = Z

X

ε

2 |∇

H

u |

2H

+ W (u) ε

≥ Z

X

p 2W (u

ε

) |∇

H

u |

H

= Z

1

0

p 2W (t) P

γ

( { u

ε

> t } ) dt .

Fix now δ > 0. From the fact that γ ( { δ ≤ u

ε

≤ 1 − δ } ) → 0 as ε → 0, it follows that, for every sequence t

ε

∈ [δ, 1 − δ], then functions χ

{uε>tε}

tend weakly to u in L

2γ

(X). For every ε > 0 let us choose t

ε

∈ [δ, 1 − δ] such that

Z

1−δ

δ

p 2W (t)P

γ

( { u

ε

> t } )dt ≥

Z

1−δ δ

p 2W (t)dt

P

γ

( { u

ε

> t

ε

} ).

Then, by Theorem 4.6 we have lim

ε→0

F

ε

(u

ε

) ≥ lim

ε→0

Z

1−δ δ

p 2W (t)dt

P

γ

( { u

ε

> t

ε

} )

Z

1−δ δ

p 2W (t)dt

F (u) . Since δ is arbitrary we get the Γ-liminf inequality.

The Γ-limsup is done similarly to the (Euclidean) finite dimensional case [29, 28]. Since F is the relaxation of F in the weak L

2γ

(X)-topology and since we can approximate sets of finite perimeter by smooth cylindrical sets by Proposition 2.6, for every u ∈ BV

γ

(X) with 0 ≤ u ≤ 1 there exists a sequence E

n

of smooth cylindrical sets with χ

En

converging weakly to u and such that P

γ

(E

n

) tends to F (u). This shows that we can restrict ourselves to smooth cylindrical sets for computing the Γ-limsup of F

ε

.

Let m ∈ N and E = Π

−1m

(E

m

), where E

m

⊂ R

m

is a smooth set with finite Gaussian perimeter, and let

d

H

(x, E) := d(Π

m

(x), E

m

)

where d(x, E

m

) is the usual distance function from E

m

in R

m

. Notice that d

H

(x, E) = min {| x − y |

H

; y ∈ E, x − y ∈ H } ,

moreover d

H

is differentiable almost everywhere with |∇

H

d

H

(x, E) |

H

= 1.

Références

Documents relatifs

In a Wiener space with a Hilbert structure there are two possible definitions of gradient, and consequently two dif- ferent notions of Sobolev spaces, functions of bounded variation

OEIGINAL: ENGLISH APPLICATION BY THE BELGIAN GOVERNMENT FOE WHO APPEOVAL OF THE DESIGNATION OF THE STANLEYVILLE LABOEATOEY, BELGIAN CONGO, AS AW INSTITUTE EEGULAKLY CARRYING

Aware of the grave consequences of substance abuse, the United Nations system, including the World Health Organization, has been deeply involved in many aspects of prevention,

For conservative dieomorphisms and expanding maps, the study of the rates of injectivity will be the opportunity to understand the local behaviour of the dis- cretizations: we

In order to study the validity of the 2.5D model for non planar geometry (Section 2), we compare numerical simulations performed on this model with predictions given by the

The fact that all fields have class number one is enough for the discussion of quadratic rings in §5; for the statements of descent in §6 concerning minimal fields of definition

A natural question about a closed negatively curved manifold M is the following: Is the space MET sec&lt;0 (M) of negatively curved met- rics on M path connected?. This problem has

First introduced by Faddeev and Kashaev [7, 9], the quantum dilogarithm G b (x) and its variants S b (x) and g b (x) play a crucial role in the study of positive representations