Home Regularity results for an optimal design problem with lower order terms
Article Open Access

Regularity results for an optimal design problem with lower order terms

  • Luca Esposito ORCID logo EMAIL logo and Lorenzo Lamberti ORCID logo
Published/Copyright: October 25, 2022

Abstract

We study the regularity of the interface for optimal energy configurations of functionals involving bulk energies with an additional perimeter penalization of the interface. Here we allow a more general structure for the energy functional in the bulk term. For a minimal configuration ( E , u ) , the Hölder continuity of u is well known. We give an estimate for the singular set of the boundary E . Namely we show that the Hausdorff dimension of the singular set is strictly smaller than n - 1 .

MSC 2010: 49Q10; 49N60; 49Q20

1 Introduction and statements

In this paper we will deal with energy functionals of the type

(1.1) ( E , u ; Ω ) = Ω [ F ( x , u , u ) + 𝟙 E G ( x , u , u ) ] 𝑑 x + P ( E , Ω ) ,

where u H 1 ( Ω ) and 𝟙 E denotes the characteristic function of a set E Ω with finite perimeter in Ω, denoted as P ( E , Ω ) . In mathematical and physical literature, the problem of finding the minimal energy configuration of a mixture of two materials in a bounded connected open set Ω n has been widely investigated (see for instance [1, 2, 11, 12, 15, 18, 19, 20, 24]). The energy functional employed in such problems involves both bulk and interface energies in order to describe a large class of phenomena in many applied sciences, such as non-linear elasticity, material sciences and image segmentation in the computer vision. A relevant case deeply studied by several authors is the following model functional:

(1.2) Ω σ E ( x ) | u | 2 𝑑 x + P ( E , Ω ) ,

where u = u 0 is prescribed on Ω and σ E ( x ) = β 𝟙 E + α 𝟙 Ω E , with 0 < α < β given constants.

In 1993 L. Ambrosio and G. Buttazzo in [2] proved that if ( E , u ) is a minimizer of the functional (1.2), then u is locally Hölder continuous in Ω and E is relatively open in Ω. At the same time and in the same volume of the same journal, F. H. Lin proved the regularity of the interface (see [20]), that is, E is regular outside a relatively closed set of vanishing n - 1 -measure. To be more precise, we define the set of regular points of E as follows:

Reg ( E ) := { x E Ω : E  is a  C 1 , γ  hypersurface in some  I ( x )  and for some  γ ( 0 , 1 ) } ,

where I ( x ) denotes a neighborhood of x. Accordingly, we define the set of singular points of E

Σ ( E ) := ( E Ω ) Reg ( E ) .

In the paper by F. H. Lin (see [20]) it was proved that n - 1 ( Σ ( E ) ) = 0 for minimal configurations of the functional (1.2). More recently, G. De Philippis and A. Figalli in [8], N. Fusco and V. Julin in [13], independently of each other and by different approaches, were able to improve this result showing that

dim ( Σ ( E ) ) n - 1 - ε ,

for some ε > 0 depending only on α , β . Regarding this dependence, it is worth noticing that in [10] it was proven that u C 0 , 1 2 + ε and the reduced boundary * E of E is a C 1 , ε -hypersurface and s ( E * E ) = 0 for all s > n - 8 , assuming that 1 α β < γ n , for some γ n > 1 depending only on the dimension.

In 1999 F.H. Lin and R.V. Kohn in [21] treated a more general quadratic bulk energy of the type (1.1) actually proving the same regularity proved by the first author for the model case (1.2). Indeed, they proved that the singular part Σ ( E ) has vanishing n - 1 -measure for minimal configurations ( E , u ) . In this paper we address the issue of improving the dimensional estimate for the singular part Σ ( E ) of optimal configurations for the class of functionals treated by F. H. Lin and R. V. Kohn. For a wide class of quadratic functionals depending also on x and u, we prove the same kind of regularity proved in the model case (1.2) by [8] and [13], namely dim ( Σ ( E ) ) n - 1 - ε for some ε > 0 . Our path to prove the aforementioned result basically follows the same strategy used in [13]. As we will point out later in more detail, our technique relies on the linearity of the Euler–Lagrange equation of the functional (1.1). For this reason we need a linear structure condition for the bulk energy. Conversely, the nonlinear case is less studied and there are few regularity results available (see [4, 5, 9, 17]).

Throughout the paper we will assume that the density energies F and G in (1.1) satisfy the following assumptions:

(1.3) F ( x , s , z ) = i , j = 1 n a i j ( x , s ) z i z j + i = 1 n a i ( x , s ) z i + a ( x , s ) ,
(1.4) G ( x , s , z ) = i , j = 1 n b i j ( x , s ) z i z j + i = 1 n b i ( x , s ) z i + b ( x , s ) .

Concerning the coefficients we assume that

a i j , b i j , a i , b i , a , b C 0 , 1 ( Ω × ) .

We will denote by L D the greatest Lipschitz constant of the data a i j , b i j , a i , b i , a , b , that is,

(1.5) | a i j | L D , | b i j | L D in  Ω × ,

and the same holds true for a i , b i , a , b .

Moreover, to ensure the existence of minimizers we assume the boundedness of the coefficients and the ellipticity of the matrices a i j and b i j ,

(1.6) ν | z | 2 a i j ( x , s ) z i z j N | z | 2 , ν | z | 2 b i j ( x , s ) z i z j N | z | 2 ,
(1.7) i = 1 n | a i ( x , s ) | + i = 1 n | b i ( x , s ) | + | a ( x , s ) | + | b ( x , s ) | L ,

where ν, N and L are three positive constants. We are interested in the regularity of minimizers of the following constrained problem.

Definition 1.

We shall denote by ((${\mathrm{P}_{\mathrm{c}}}$)) the constrained problem

(${\mathrm{P}_{\mathrm{c}}}$) min E 𝒜 ( Ω ) v u 0 + H 0 1 ( Ω ) { ( E , v ; Ω ) : | E | = d } ,

where u 0 H 1 ( Ω ) , 0 < d < | Ω | are given and 𝒜 ( Ω ) is the class of all subsets of Ω with finite perimeter in Ω.

The problem of handling with the constraint | E | = d is overtaken using an argument introduced in [10], ensuring that every minimizer of the constrained problem ((${\mathrm{P}_{\mathrm{c}}}$)) is also a minimizer of a penalized functional of the type

Λ ( E , v ; Ω ) = ( E , v ; Ω ) + Λ | | E | - d | ,

for some suitable Λ > 0 (see Theorem 2 below). Therefore, we give in addition the following definition.

Definition 2.

We shall denote by (($\mathrm{P}$)) the penalized problem

($\mathrm{P}$) min E 𝒜 ( Ω ) v u 0 + H 0 1 ( Ω ) Λ ( E , v ; Ω ) ,

where u 0 H 1 ( Ω ) is fixed and 𝒜 ( Ω ) is the same class defined in Definition 1.

From the point of view of regularity, the extra term Λ | | E | - | F | | is a higher order negligible perturbation. The main result of the paper is stated in the following theorem.

Theorem 1.

Let ( E , u ) be a minimizer of problem (($\mathrm{P}$)), under assumptions (1.3)–(1.7). Then:

  1. there exists a relatively open set Γ E such that Γ is a C 1 , μ hypersurface for all 0 < μ < 1 2 ,

  2. there exists ε > 0 depending on ν , N , L , n , such that

    n - 1 - ε ( ( E Γ ) Ω ) = 0 .

For the reader’s convenience the paper is structured in sections which reflect the proof strategy. Section 2 collects known results and preliminary definitions. As in the case of minimizers of the Mumford–Shah functional, the proof of regularity is based on the study of the interplay between the perimeter and the bulk energy (see [3, 20]). We point out that the Hölder exponent 1 2 is critical for solutions u of either (($\mathrm{P}$)) or ((${\mathrm{P}_{\mathrm{c}}}$)), in the sense that, whenever u C 0 , 1 2 , under appropriate scaling, the bulk term locally has the same dimension n - 1 as the perimeter term. In this regard, our starting point is to prove suitable energy decay estimates for the bulk energy. These estimates are presented in Section 3. The key point of this approach is contained in Lemma 8, where it is proved that the bulk energy decays faster than ρ n - 1 , that is, for any δ > 0 ,

(1.8) B ρ ( x 0 ) | u | 2 𝑑 x C ρ n - δ ,

either in the case that

min { | E B ρ ( x 0 ) | , | B ρ ( x 0 ) E | } < ε 0 | B ρ ( x 0 ) | ,

or in the case that there exists an half-space H such that

| ( Δ E H ) B ρ ( x 0 ) | ε 0 | B ρ ( x 0 ) | ,

for some ε 0 > 0 . The latter case is the hardest one to handle because it relies on the regularity properties of solutions of a transmission problem which we study in Section 3.1. Let us notice that, for any given E Ω , local minimizers u of the functional

(1.9) Ω [ F ( x , u , u ) + 𝟙 E G ( x , u , u ) ] 𝑑 x

are Hölder continuous, u C loc 0 , α ( Ω ) , but the needed bound α > 1 2 cannot be expected in the general case without any information on the set E. In Section 3.1 we prove that minimizers of the functional (1.9) are in C 0 , α for every α ( 0 , 1 ) , in the case E is an half-space. In this context the linearity of the equation strongly comes into play ensuring that the derivatives of the Euler–Lagrange equation are again solutions of the same equation. For the proof in Section 3 we readapt a technique depicted in the book [3] in the context of the Mumford–Shah functional and recently used in a paper by E. Mukoseeva and G. Vescovo, [23]. Once obtained the estimates of Section 3, in Section 4 we are in a position to prove that, if in a ball B ρ ( x 0 ) the perimeter of E is sufficiently small, then the total energy

B r ( x 0 ) | u | 2 𝑑 x + P ( E , B r ( x 0 ) ) , 0 < r < ρ ,

decays as r n (see Lemma 9). Making use of the latter energy density estimate we are in a position to deduce in Section 4 the density lower bound for the perimeter of E as well. In the subsequent sections the proof strategy follows the path traced from the regularity theory for perimeter minimizers. In Section 5 it is proved the compactness for sequences of minimizers which more or less follows in a standard way from the density lower bound. Section 6 is devoted to proving some additional consequences of the density lower bound which involve the excess

𝐞 ( x , r ) = inf ν 𝕊 n - 1 𝐞 ( x , r , ν ) := inf ν 𝕊 n - 1 1 r n - 1 E B r ( x ) | ν E ( y ) - ν | 2 2 𝑑 n - 1 ( y ) .

Indeed, we prove the height bound lemma and the Lipschitz approximation theorem. In this section we also compute the Euler–Lagrange equation for ( E , u ) involving the variation of the set E. Section 7 is devoted to prove the excess improvement which follows from the fact that whenever the excess 𝐞 ( x , r ) goes to zero, for r 0 , the Dirichlet integral B ρ ( x 0 ) | u | 2 𝑑 x decays as in (1.8). In Section 8 we provide the proof of Theorem 1 that is a consequence of the excess improvement proved before.

2 Preliminary notations and definitions

In the rest of the paper we will write ξ , η for the inner product of vectors ξ , η n , and | ξ | := ξ , ξ 1 2 will denote the corresponding Euclidean norm. We will denote by ω n the Lebesgue measure of the unit ball in n and by B R ( x 0 ) the ball with center x 0 n and radius R > 0 . If x 0 = 0 , we will simply write B R . If u is integrable in B R ( x 0 ) , we set

u R = 1 ω n R n B R ( x 0 ) u 𝑑 x = B R ( x 0 ) u 𝑑 x .

For any μ 0 we define the Morrey space L 2 , μ ( Ω ) as

(2.1) L 2 , μ ( Ω ) := { u L 2 ( Ω ) : sup x 0 Ω , r > 0 r - μ Ω B r ( x 0 ) | u | 2 𝑑 x < } .

The following definition is standard.

Definition 3.

Let v H loc 1 ( Ω ) and assume that E Ω is fixed. We define the functional E as

E ( w , Ω ) := ( E , w ; Ω ) for all  w H 1 ( Ω ) .

Furthermore, we say that v is a local minimizer of the integral functional E if and only if

E ( v ; B R ( x 0 ) ) = min w v + H 0 1 ( B R ( x 0 ) ) E ( w ; B R ( x 0 ) ) ,

for all B R ( x 0 ) Ω .

It is clear that any minimizer u of problem ((${\mathrm{P}_{\mathrm{c}}}$)) is a local minimizer of the functional (1.1) and therefore satisfies the Euler–Lagrange equation

i = 1 n x i [ F z i ( x , u , u ) + 𝟙 E ( x ) G z i ( x , u , u ) ] = F u ( x , u , u ) + 𝟙 E ( x ) G u ( x , u , u ) .

It is worth mentioning that for a quadratic integrand F ( x , s , z ) of the type given in (1.3) the following growth condition can be immediately deduced from assumptions (1.6) and (1.7):

(2.2) ν 2 | z | 2 - L 2 ν F ( x , s , z ) ( N + 1 ) | z | 2 + L ( L + 1 ) for all  x Ω , all  s  and all  z n .

Lemma 1.

Let Z ( t ) be a bounded non-negative function in the interval [ ρ , R ] and assume that for ρ t < s R we have

Z ( t ) θ Z ( s ) + A ( s - t ) 2 + B ,

with A , B 0 and 0 θ < 1 . Then

Z ( ρ ) c [ A ( R - ρ ) 2 + B ] ,

for some c = c ( θ ) .

Proof.

The proof of this lemma is standard and can be found in [14]. Inspecting the proof in [14] one can also obtain an explicit expression of the constant c ( θ ) . An admissible value for c ( θ ) , but not the best one, is c ( θ ) = 1 1 - θ 1 3 ) 3 . ∎

The next lemma can be found in [3, Lemma 7.54].

Lemma 2.

Let f : ( 0 , a ] [ 0 , ) be an increasing function such that

f ( ρ ) A [ ( ρ R ) p + R s ] f ( R ) + B R q whenever  0 < ρ < R a ,

for some constants A , B 0 , 0 < q < p , s > 0 . Then there exist R 0 = R 0 ( p , q , s , A ) and c = c ( p , q , A ) such that

f ( ρ ) c ( ρ R ) q f ( R ) + c B ρ q whenever  0 < ρ < R min { R 0 , a } .

2.1 From constrained to penalized problem

The next theorem allows us to overcome the difficulty of handling with the constraint | E | = d . Indeed, it can be proved that every minimizer of the constrained problem ((${\mathrm{P}_{\mathrm{c}}}$)) is also a minimizer of a suitable unconstrained problem with a volume penalization of the type given in (($\mathrm{P}$)).

Theorem 2.

There exists Λ 0 > 0 such that if ( E , u ) is a minimizer of the functional

(2.3) Λ ( A , w ) = Ω [ F ( x , w , w ) + 𝟙 A G ( x , w , w ) d x ] 𝑑 x + P ( A , Ω ) + Λ | | A | - d | ,

for some Λ Λ 0 , among all configurations ( A , w ) such that w = u 0 on Ω , then | E | = d and ( E , u ) is a minimizer of problem ((${\mathrm{P}_{\mathrm{c}}}$)). Conversely, if ( E , u ) is a minimizer of problem ((${\mathrm{P}_{\mathrm{c}}}$)), then it is a minimizer of (2.3), for all Λ Λ 0 .

Proof.

The proof can be carried out as in [10, Theorem 1]. For the reader’s convenience we give here its sketch, emphasizing main ideas and minor differences with respect to the case treated in [10].

The first part of the theorem can be proved by contradiction. Assume that there exist a sequence ( λ h ) h such that λ h as h and a sequence of configurations ( E h , u h ) minimizing λ h and such that u h = u 0 on Ω and | E h | d for all h . Let us choose now an arbitrary fixed E 0 Ω with finite perimeter such that | E 0 | = d . Let us point out that

(2.4) λ h ( E h , u h ) ( E 0 , u 0 ) := Θ .

Without loss of generality we may assume that | E h | < d . Indeed, the case | E h | > d can be treated in the same way considering the complement of E h in Ω. Our aim is to show that for h sufficiently large, there exists a configuration ( E ~ h , u ~ h ) such that λ h ( E ~ h , u ~ h ) < λ h ( E h , u h ) , thus proving the result by contradiction.

By condition (2.4), it follows that the sequence { u h } is bounded in H 1 ( Ω ) , the perimeters of the sets E h in Ω are bounded and | E h | d . Therefore, possibly extracting a not relabelled subsequence, we may assume that there exists a configuration ( E , u ) such that u h u weakly in H 1 ( Ω ) , 𝟙 E h 𝟙 E a.e. in Ω, where the set E is of finite perimeter in Ω and | E | = d . The couple ( E , u ) will be used as reference configuration for the definition of ( E ~ h , u ~ h ) .

Step 1: Construction of ( E ~ h , u ~ h ) . Proceeding exactly as in [10], we take a point x * E Ω and observe that the sets E r = E - x r converge locally in measure to the half-space H = { z , ν E ( x ) < 0 } , i.e., 𝟙 E r 𝟙 H in L loc 1 ( n ) , where ν E ( x ) is the generalized exterior normal to E at x (see [3, Definition 3.54]). Let y B 1 ( 0 ) H be the point y = ν E ( x ) 2 . Given ε (that will be chosen in Step 4), since 𝟙 E r 𝟙 H in L 1 ( B 1 ( 0 ) ) there exists 0 < r < 1 such that

| E r B 1 / 2 ( y ) | < ε , | E r B 1 ( y ) | | E r B 1 / 2 ( 0 ) | > ω n 2 n + 2 ,

where ω n denotes the measure of the unit ball of n . Then if we define x r = x + r y Ω , we have

| E B r / 2 ( x r ) | < ε r n , | E B r ( x r ) | > ω n r n 2 n + 2 .

Let us assume, without loss of generality, that x r = 0 . From the convergence of E h to E we have that for all h sufficiently large

(2.5) | E h B r / 2 | < ε r n , | E h B r | > ω n r n 2 n + 2 .

Let us now define the following bi-Lipschitz function used in [10] which maps B r into itself:

(2.6) Φ ( x ) = { ( 1 - σ h ( 2 n - 1 ) ) x if  | x | < r 2 , x + σ h ( - r n | x | n ) x if  r 2 | x | < r , x if  | x | r ,

for some 0 < σ h < 1 2 n sufficiently small in such a way that, setting

E ~ h = Φ ( E h ) , u ~ h = u h Φ - 1 ,

we have | E ~ h | < d . We obtain

(2.7)

λ h ( E h , u h ) - λ h ( E ~ h , u ~ h ) = [ B r [ F ( x , u h , u h ) + 𝟙 E h G ( x , u h , u h ) ] d x
- B r [ F ( x , u ~ h , u ~ h ) + 𝟙 E ~ h G ( x , u ~ h , u ~ h ) ] d y ]
+ [ P ( E h , B ¯ r ) - P ( E ~ h , B ¯ r ) ] + λ h ( | E ~ h | - | E h | )
= : I 1 , h + I 2 , h + I 3 , h .

Step 2: Estimate of I 1 , h . In order to estimate the bulk energy I 1 , h we write down some preliminary estimates for the map Φ that can be obtained by direct computation (see [10] or [4] for the explicit calculation). We just sketch the main point. First observe that, for | x | < r 2 , Φ is simply a homothety and all the estimates that we need are very easy to obtain. Conversely, for r 2 < | x | < r we have

(2.8) Φ i x j ( x ) = ( 1 + σ h - σ h r n | x | n ) δ i j + n σ h r n x i x j | x | n + 2 .

It is clear from this expression that, since σ h is small, Φ is a small perturbation of the identity in the sense that

(2.9) | z - z Φ ( y ) | C 1 ( n ) σ h | z | for all  y , z n .

It is not difficult to find out also that

(2.10) Φ - 1 ( Φ ( x ) ) ( 1 - ( 2 n - 1 ) σ h ) - 1 1 + 2 n n σ h for all  x B r .

Concerning J Φ , the Jacobian of Φ, from (2.8) we deduce

J Φ ( x ) = ( 1 + σ h + ( n - 1 ) σ h r n | x | n ) ( 1 + σ h - σ h r n | x | n ) n - 1 .

Using the fact that r 2 < | x | < r , we can estimate (see also [4, Section 3]):

J Φ ( x ) ( 1 + σ h + ( n - 1 ) σ h r n | x | n ) ( 1 + σ h - ( n - 1 ) σ h r n | x | n )
1 + 2 σ h - ( 4 n ( n - 1 ) 2 - 1 ) σ h 2 > 1 + σ h ,

provided that we choose

σ h < 1 4 n ( n - 1 ) 2 - 1 .

Summarizing we gain the following inequalities for the Jacobian of Φ:

(2.11) 1 + σ h J Φ ( x ) 1 + 2 n n σ h for all  x B r B r / 2 .

Now we can perform the change of variables y = Φ ( x ) and, observing that 𝟙 E ~ h ( Φ ( x ) ) = 𝟙 E h ( x ) , we get

I 1 , h = B r [ F ( x , u h , u h ) - J Φ ( x ) F ( Φ ( x ) , u h ( x ) , u h ( x ) Φ - 1 ( Φ ( x ) ) ) ] 𝑑 x
+ B r E h [ G ( x , u h , u h ) - J Φ ( x ) G ( Φ ( x ) , u h ( x ) , u h ( x ) Φ - 1 ( Φ ( x ) ) ) ] 𝑑 x
= : J 1 , h + J 2 , h .

The two terms J 1 , h and J 2 , h , involving F and G in B r and B r E h respectively, can be treated in the same way. Therefore we just perform the calculation for J 1 , h . To make the argument more clear, since we shall use the structure conditions (1.3) and (1.4), we introduce the following notations. A 2 ( x , s ) denotes the quadratic form and A 1 ( x , s ) denotes the linear form defined as follows:

A 2 ( x , s ) [ z ] := a i j ( x , s ) z i z j , A 1 ( x , s ) [ z ] := a i ( x , s ) z i

for any z n . Analogously we set A 0 ( x , s ) = a ( x , s ) . Accordingly, we can write down

(2.12) J 1 , h = B r { A 2 ( x , u h ( x ) ) [ u h ( x ) ] - A 2 ( Φ ( x ) , u h ( x ) ) [ u h ( x ) Φ - 1 ( Φ ( x ) ) ] J Φ ( x ) } 𝑑 x + B r { A 1 ( x , u h ( x ) ) [ u h ( x ) ] - A 1 ( Φ ( x ) , u h ( x ) ) [ u h ( x ) Φ - 1 ( Φ ( x ) ) ] J Φ ( x ) } 𝑑 x + B r { A 0 ( x , u h ( x ) ) - A 0 ( Φ ( x ) , u h ( x ) ) J Φ ( x ) } 𝑑 x .

We proceed estimating the first difference in the previous equality, being the other similar and indeed easier to handle:

B r { A 2 ( x , u h ( x ) ) [ u h ( x ) ] - A 2 ( Φ ( x ) , u h ( x ) ) [ u h ( x ) Φ - 1 ( Φ ( x ) ) ] J Φ ( x ) } 𝑑 x
= B r { A 2 ( Φ ( x ) , u h ( x ) ) [ u h ( x ) ] - A 2 ( Φ ( x ) , u h ( x ) ) [ u h ( x ) Φ - 1 ( Φ ( x ) ) ] J Φ ( x ) } 𝑑 x
+ B r { A 2 ( x , u h ( x ) ) [ u h ( x ) ] - A 2 ( Φ ( x ) , u h ( x ) ) [ u h ( x ) ] } d x = : H 1 , h + H 2 , h .

The first term H 1 , h can be estimated observing that, as a consequence of (1.6), we have

| A 2 [ ξ ] - A 2 [ η ] | N | ξ + η | | ξ - η | for all  ξ , η n .

If we apply the last inequality to the vectors

ξ := u h ( x ) , η := J Φ ( x ) [ u h ( x ) Φ - 1 ( Φ ( x ) ) ] ,

we are led to estimate | ξ - η | . We start observing that, being J Φ ( x ) = ( 1 - σ h ( 2 n - 1 ) ) n for | x | < r 2 , by also using (2.11), we deduce

| J Φ ( x ) - 1 | < C ( n ) σ h for all  x n .

Therefore we have

| J Φ ξ - ξ | C ( n ) σ h | ξ | .

In addition, choosing z = ξ Φ - 1 ( Φ ( x ) ) in (2.9) and using also (2.10), we can deduce

| ξ Φ - 1 ( Φ ( x ) ) - ξ | σ h C 1 ( n ) | ξ Φ - 1 ( Φ ( x ) ) | σ h | ξ | C 1 ( n ) Φ - 1 ( Φ ( x ) ) n 2 n C 1 ( n ) σ h | ξ | .

Summarizing we finally get

| ξ - η | σ h C ( n ) | u h ( x ) | , | ξ + η | C ( n ) | u h ( x ) | ,

for some constant C = C ( n ) > 0 . From the previous estimates we deduce that

(2.13) | H 1 , h | σ h N C 2 ( n ) B r | u h ( x ) | 2 𝑑 x σ h N C 2 ( n ) Θ ,

where Θ is defined in (2.4). The second term H 2 , h can be estimated using the Lipschitz assumption of a i j and observing that | x - Φ ( x ) | σ h r 2 n . Therefore we deduce that

(2.14) | H 2 , h | σ h r 2 n L D B r | u h ( x ) | 2 𝑑 x σ h C ( n , L D ) Θ .

In conclusion, since the other terms in (2.12) can be estimated in the same way, collecting estimates (2.13) and (2.14), we get

| J 1 , h | σ h C ( n , N , L D ) Θ .

Since the same estimate holds true for J 2 , h , we conclude that

(2.15) I 1 , h - σ h C 2 ( n , N , L D ) Θ ,

for some constant C 2 = C 2 ( n , N , L D ) > 0 .

Step 3: Estimate of I 2 , h . In order to estimate I 2 , h , we can use the area formula for maps between rectifiable sets. If we denote by T h , x the tangential gradient of Φ along the approximate tangent space to * E h in x and T h , x * is the adjoint of the map T h , x , the ( n - 1 ) -dimensional jacobian of T h , x is given by

J n - 1 T h , x = det ( T h , x * T h , x ) .

Thereafter we can estimate

(2.16) J n - 1 T h , x 1 + σ h + 2 n ( n - 1 ) σ h .

We address the reader to [10] where explicit calculations are given. In order to estimate I 2 , h , we use the area formula for maps between rectifiable sets ([3, Theorem 2.91]), thus getting

I 2 , h = P ( E h , B ¯ r ) - P ( E ~ h , B ¯ r )
= * E h B ¯ r 𝑑 n - 1 - * E h B ¯ r J n - 1 T h , x 𝑑 n - 1
= * E h B ¯ r B r / 2 ( 1 - J n - 1 T h , x ) 𝑑 n - 1 + * E h B r / 2 ( 1 - J n - 1 T h , x ) 𝑑 n - 1 .

Notice that the last integral in the above formula is non-negative since Φ is a contraction in B r / 2 , hence J n - 1 T h , x < 1 in B r / 2 , while from (2.16) we have

* E h B ¯ r B r / 2 ( 1 - J n - 1 T h , x ) 𝑑 n - 1 - 2 n n P ( E h , B ¯ r ) σ h - 2 n n Θ σ h ,

thus concluding that

(2.17) I 2 , h - 2 n n Θ σ h .

Step 4: Estimate of I 3 , h . To estimate I 3 , h , we recall (2.5), (2.6) and (2.11), thus getting

I 3 , h = λ h E h B r B r / 2 ( J Φ ( x ) - 1 ) 𝑑 x + λ h E h B r / 2 ( J Φ ( x ) - 1 ) 𝑑 x
λ h ( ω n 2 n + 2 - ε ) σ h r n - λ h [ 1 - ( 1 - ( 2 n - 1 ) σ h ) n ] ε r n
λ h σ h r n [ ω n 2 n + 2 - ε - ( 2 n - 1 ) n ε ] .

Therefore, if we choose 0 < ε < ε ( n ) , we have that

I 3 , h λ h C 3 ( n ) σ h r n ,

for some positive C 3 = C 3 ( n ) . From this inequality, recalling (2.7), (2.15) and (2.17), we obtain

λ h ( E h , u h ) - λ h ( E ~ h , u ~ h ) σ h ( λ h C 3 r n - Θ ( C 2 ( n , N , L D ) + 2 n n ) ) > 0

if λ h is sufficiently large. This contradicts the minimality of ( E h , u h ) , thus concluding the proof. ∎

The previous theorem motivates the following definition.

Definition 4 (Λ-minimizers).

The energy pair ( E , u ) is a Λ-minimizer in Ω of the functional , defined in (1.1), if and only if for every B r ( x 0 ) Ω it holds

( E , u ; B r ( x 0 ) ) ( F , v ; B r ( x 0 ) ) + Λ | Δ F E |

whenever ( F , v ) is an admissible test pair, namely, F is a set of finite perimeter with F Δ E B r ( x 0 ) and v - u H 0 1 ( B r ( x 0 ) ) .

3 Decay of the bulk energy

In the first part of this section we collect some preliminary results concerning decay estimates for local minimizers u of the functional (1.1) when E is fixed. We start quoting higher integrability results both for local minimizers of functional (1.1) and for comparison functions that we will use later in the paper. It is worth mentioning that the following lemmata can be applied in general to minimizers of integral functionals of the type

(3.1) ( u ; Ω ) := Ω F ( x , u , u ) 𝑑 x ,

assuming that the energy density only satisfies the structure condition (1.3) and the growth conditions (1.6) and (1.7), without assuming any continuity on the coefficients. Therefore, functionals of the type (1.1) belong to this class and in addition the involved estimates only depend on the constants appearing in (1.6) and (1.7) but do not depend on E accordingly.

Lemma 3.

Let u H 1 ( Ω ) be a local minimizer of the functional H defined in (3.1), where F satisfies the structure condition (1.3) and the growth conditions (1.6) and (1.7). Then for every B 2 R ( x 0 ) Ω it holds

(3.2) B R ( x 0 ) | u | 2 𝑑 x C 1 [ ( B 2 R ( x 0 ) | u | 2 m 𝑑 x ) 1 m + 1 ] ,

where m = n n + 2 and C 1 = C 1 ( n , ν , N , L ) is a positive constant.

Proof.

Without loss of generality we may assume that x 0 = 0 . Let R < t < s < 2 R and choose η C 0 ( B s ) such that η 1 in B t and | η | 2 s - t . We choose a test function v = u - ϕ , where ϕ = η ( u - u s ) and u s denotes the average of u in B s ,

u s = B s u 𝑑 x .

Testing the minimality of u with v and using growth condition (2.2) we deduce that

ν 2 B s [ | u | 2 - 2 L 2 ν 2 ] 𝑑 x ( u ; B s ) ( v ; B s )
2 B s [ ( N + 1 ) | u ( 1 - η ) - η ( u - u s ) | 2 + L ( L + 1 ) ] 𝑑 x
4 ( N + 1 ) B s B t | u | 2 𝑑 x + 4 ( N + 1 ) B s | u - u s | 2 | η | 2 𝑑 x + 2 B s L ( L + 1 ) 𝑑 x .

Adding to both sides 4 ( N + 1 ) B t | u | 2 𝑑 x , we deduce

[ 4 ( N + 1 ) + ν 2 ] B t | u | 2 𝑑 x 4 ( N + 1 ) B s | u | 2 𝑑 x + 4 ( N + 1 ) B s | u - u s | 2 | η | 2 𝑑 x + B s C ( L , ν ) 𝑑 x .

Thus we get

B t | u | 2 𝑑 x θ B s | u | 2 𝑑 x + C ( n , ν , N , L ) ( s - t ) 2 B s | u - u s | 2 𝑑 x + C ( n , ν , N , L ) ,

where θ = 4 ( N + 1 ) 4 ( N + 1 ) + ν / 2 < 1 . Using Sobolev–Poincaré’s inequality

B s | u - u s | 2 𝑑 x C ( n ) ( B s | u | 2 m 𝑑 x ) 1 m C ( n ) ( B 2 R | u | 2 m 𝑑 x ) 1 m ,

with m = n n + 2 , we eventually obtain

B t | u | 2 𝑑 x θ B s | u | 2 𝑑 x + C ( n , ν , N , L ) ( s - t ) 2 ( B 2 R | u | 2 m 𝑑 x ) 1 m + C ( n , ν , N , L ) .

Iterating the previous estimate using Lemma 1, we deduce that there exists a constant C = C ( θ ) = C ( ν , N ) > 0 such that

B R | u | 2 𝑑 x C ( ν , N ) [ C ( n , ν , N , L ) R 2 ( B 2 R | u | 2 m 𝑑 x ) 1 m + C ( n , ν , N , L ) ] .

We get the thesis if we divide by R n . ∎

Remark 1.

We observe that the reverse Hölder inequality stated in the previous lemma can be also proved exactly in the same way replacing the balls with the cubes. The reverse Hölder inequality written on cubes is the suitable version in order to apply Calderòn–Zygmund decomposition and Gehring’s lemma (see [14, Proposition 6.1]), thus obtaining the higher integrability estimate on cubes. Finally, the higher integrability estimate on balls stated below, which is suitable in our setting, can be deduced by a covering argument (see [13, Appendix]).

Lemma 4.

Let u H 1 ( Ω ) be a local minimizer of the functional H defined in (3.1), where F satisfies the structure condition (1.3) and the growth conditions (1.6) and (1.7). There exists s = s ( n , ν , N , L ) > 1 such that, for every B 2 R ( x 0 ) Ω , it holds

B R ( x 0 ) | u | 2 s 𝑑 x C 2 ( B 2 R ( x 0 ) ( 1 + | u | 2 ) 𝑑 x ) s ,

where C 2 = C 2 ( n , ν , N , L ) is a positive constant.

In the next subsection we will prove some energy density estimates by using a standard comparison argument. For this purpose we will need a reverse Hölder inequality for the comparison function defined below.

Definition 5 (Comparison function).

Let u H 1 ( Ω ) be a local minimizer of the functional defined in (3.1) and B 2 R Ω . We shall denote by v the solution of the following problem:

(3.3) v := arg min w u + H 0 1 ( B R ) B R F ~ ( x , w ) 𝑑 x ,

where F ~ ( x , z ) := F ( x , u ( x ) , z ) satisfies the structure condition (1.3) and the growth conditions (1.6)–(1.7).

For the comparison function v defined in (3.3) we can state the following reverse Hölder inequality up to the boundary of B R .

Lemma 5.

Let u H 1 ( Ω ) be a local minimizer of the functional H defined in (3.1), where F satisfies the structure condition (1.3) and the growth conditions (1.6) and (1.7). Let v be the comparison function defined above and B 2 R Ω . Let us consider the following extension of v:

V ( x ) := { v ( x ) for  x B R , u ( x ) for  x Ω B R .

Let B ρ ( x 0 ) B 2 R with x 0 B R and ρ < R 2 . Then

(3.4) B ρ ( x 0 ) | V | 2 𝑑 x C 3 [ ( B 2 ρ ( x 0 ) | V | 2 m 𝑑 x ) 1 m + ( B 2 ρ ( x 0 ) | u | 2 m ) 1 m + 1 ] ,

where m = n n + 2 and C 3 = C 3 ( n , ν , N , L ) is a positive constant.

Proof.

Let x 0 B R and ρ s < t r < 2 ρ < R , where r = 3 2 ρ ; then the following alternatives hold:

  1. B r ( x 0 ) B R ,

  2. B ¯ r ( x 0 ) ( Ω B R ) .

In case (i) we can proceed exactly as in Lemma 3 to get the desired estimate. Let us then consider case (ii) which is slightly different. Choose η C 0 ( B t ( x 0 ) ) , such that 0 η 1 , η 1 in B s ( x 0 ) and | η | 2 t - s . Now we can use the function φ := η ( V - u ) to test the minimality of v with the aim of estimating the difference B s ( x 0 ) | ( V - u ) | 2 . In the following it would be useful to keep in mind that φ = ( V - u ) in B s ( x 0 ) , which is the quantity we are interested to estimate. In order to simplify the notation, let us denote

a ~ i j ( x ) := a i j ( x , u ( x ) ) , a ~ i ( x ) := a i ( x , u ( x ) ) , a ~ ( x ) := a ( x , u ( x ) ) ,

and

~ ( w ; A ) := A F ~ ( x , w ) 𝑑 x .

We start comparing the energy of φ and v inside a generic set A B R . We have

~ ( φ ; A ) - ~ ( v ; A ) = A a ~ i j ( x ) i φ j φ d x + A a ~ i ( x ) i φ d x + A a ~ ( x ) 𝑑 x
- A a ~ i j ( x ) i v j v d x - A a ~ i ( x ) i v d x - A a ~ ( x ) 𝑑 x
= A a ~ i j ( x ) i ( φ - v ) j ( φ - v ) d x + 2 A a ~ i j ( x ) i v j ( φ - v ) d x
+ A a ~ i ( x ) i ( φ - v ) d x .

Using the growth conditions (1.6) and (1.7) and Young’s inequality, we deduce that

(3.5) ~ ( φ ; A ) - ~ ( v ; A ) ( N + N 2 + 1 2 ) A | ( v - φ ) | 2 𝑑 x + A | v | 2 𝑑 x + A L 2 2 𝑑 x .

Recalling the growth condition (2.2), we estimate

A | v | 2 𝑑 x 2 ν A F ~ ( x , v ) 𝑑 x + A 2 L 2 ν 2 𝑑 x .

We can conclude from (3.5) that

(3.6) ~ ( φ ; A ) ( 1 + 2 ν ) ~ ( v ; A ) + ( N + N 2 + 1 2 ) A | ( v - φ ) | 2 𝑑 x + A ( L 2 2 + 2 L 2 ν 2 ) 𝑑 x .

We compute now the previous integrals on the set B t ( x 0 ) B R . We use again the growth condition (2.2) and the minimality of v with respect to v - φ in order to estimate further on the right-hand side of the previous inequality

~ ( v ; B t ( x 0 ) B R ) ~ ( ( v - φ ) ; B t ( x 0 ) B R ) ( N + 1 ) B t ( x 0 ) B R | ( v - φ ) | 2 𝑑 x + B t ( x 0 ) B R L ( L + 1 ) 𝑑 x .

Finally, we can resume (3.6) to conclude this first part concerning the energy estimate of φ, using again (2.2),

ν 2 B t ( x 0 ) ( | φ | 2 - 2 L 2 ν 2 ) 𝑑 x B t ( x 0 ) B R F ~ ( x , φ ) 𝑑 x C ( ν , N , L ) B t ( x 0 ) ( | ( V - φ ) | 2 d x + 1 ) 𝑑 x .

We can summarize the previous estimate as follows:

(3.7) ν 2 B t ( x 0 ) | φ | 2 𝑑 x C ( ν , N , L ) B t ( x 0 ) ( | ( V - φ ) | 2 + 1 ) 𝑑 x .

Now we observe that | ( V - φ ) | | u | + ( 1 - η ) | ( V - u ) | + 2 ( t - s ) | V - u | ; then by (3.7) we deduce

(3.8)

B s ( x 0 ) | ( V - u ) | 2 𝑑 x C ( ν , N , L ) B t ( x 0 ) B s ( x 0 ) | ( V - u ) | 2 𝑑 x
+ C ( ν , N , L ) ( t - s ) 2 B t ( x 0 ) | V - u | 2 𝑑 x + C ( ν , N , L ) B t ( x 0 ) ( | u | 2 + 1 ) 𝑑 x .

Now we use the “hole-filling” technique adding C ( ν , N , L ) B s ( x 0 ) | ( V - u ) | 2 𝑑 x to both sides of (3.8) getting

B s ( x 0 ) | ( V - u ) | 2 𝑑 x θ B t ( x 0 ) | ( V - u ) | 2 𝑑 x + C ( ν , N , L ) [ 1 ( t - s ) 2 B t ( x 0 ) | V - u | 2 𝑑 x + B t ( x 0 ) ( | u | 2 + 1 ) 𝑑 x ] ,

where θ = C ( ν , N , L ) ( C ( ν , N , L ) + 1 ) . Using Lemma 1, we obtain

B ρ ( x 0 ) | ( V - u ) | 2 𝑑 x C ( ν , N , L ) ( r - ρ ) 2 B r ( x 0 ) | V - u | 2 𝑑 x + C ( ν , N , L ) B r ( x 0 ) ( 1 + | u | 2 ) 𝑑 x .

Therefore, being r = 3 2 ρ and by condition (ii), we have

| B 2 ρ ( x 0 ) B R | C | B ρ ( x 0 ) | ,

for some universal constant C = C ( n ) > 0 . We can now use Sobolev-Poincaré’s inequality for functions vanishing on a set of positive measure (see [14, inequality (3.29)]) to deduce

B ρ ( x 0 ) | ( V - u ) | 2 d x C ( n , ν , N , L ) [ ( B 2 ρ ( x 0 ) ( | ( V - u ) | 2 m d x ) 1 m + B 2 ρ ( x 0 ) ( 1 + | u | 2 ) d x ]

Finally, we can apply reverse Hölder inequality (3.2) for u in the last estimate to get (3.4). ∎

Reasoning in a similar way as above, the higher integrability for v can be obtained by means of Gehring’s lemma (see [14, Proposition 6.1]).

Lemma 6.

Let u H 1 ( Ω ) be a local minimizer of the functional H defined in (3.1), where F satisfies the structure condition (1.3) and the growth conditions (1.6) and (1.7). Let v H 1 ( B R ( x 0 ) ) be the comparison function defined in (3.3). Denoting by s = s ( n , ν , N , L ) > 1 the same exponent given in Lemma 4, it holds

B R ( x 0 ) | v | 2 s 𝑑 x C 4 ( B 2 R ( x 0 ) ( 1 + | u | 2 ) 𝑑 x ) s ,

where C 4 = C 4 ( n , ν , N , L ) is a positive constant.

3.1 A decay estimate for elastic minima

In this subsection we prove a decay estimate for elastic minima that will be crucial for the proof strategy. Indeed, we show that if ( E , u ) is a Λ-minimizer of the functional defined in (1.1) and x 0 is a point in Ω, where either the density of E is close to 0 or 1, or the set E is asymptotically close to a hyperplane, then for ρ sufficiently small we have

B ρ ( x 0 ) | u | 2 𝑑 x C ρ n - δ ,

for any δ > 0 . A preliminary result we want to mention, which will be used later, provides an upper bound for . It is rather standard and is related to the threshold Hölder exponent 1 2 of the function u, when ( E , u ) is either a solution of the constrained problem ((${\mathrm{P}_{\mathrm{c}}}$)) or a solution of the penalized problem (($\mathrm{P}$)) defined in Section 1. For the proof we address the reader to [21, Lemma 2.3] and [13]. A detailed proof in the case of costrained problems and for functionals satisfying general p-polynomial growth is contained in [4].

Theorem 3.

Let ( E , u ) be a Λ-minimizer of the functional F defined in (1.1). For every open set U Ω there exists a constant C 5 = C 5 ( U , u L 2 ( Ω ) ) > 0 such that for every B r ( x 0 ) U it holds

( E , u ; B r ( x 0 ) ) C 5 r n - 1 .

As a consequence of Theorem 3, using Poincaré’s inequality and the characterization of Campanato spaces (see for example [14, Theorem 2.9]), we can infer that u C 0 , 1 2 . We deduce the following remark.

Remark 2.

Let ( E , u ) be a Λ-minimizer of the functional defined in (1.1). For every open set U Ω there exists a constant C = C ( U , u L 2 ( Ω ) ) > 0 such that

(3.9) sup x , y U | u ( x ) - u ( y ) | | x - y | 1 2 C u L 2 ( Ω ) .

Notation 4.

In the sequel E Ω will denote any given subset of Ω with finite perimeter. We denote by u E , or simply by u if no confusion arises, any local minimizers of the functional E ( v ; Ω ) .

  1. If x n , we write x = ( x , x n ) , where x n - 1 and x n . Accordingly, we denote = ( x 1 , , x n - 1 ) the gradient with respect to the first n - 1 components.

  2. We will denote H = { x : x n > 0 } .

In order to prove the main lemma of this section we introduce the following preliminary result. For the reader’s convenience we give here a sketch of the proof, which can be found in [23]. Actually we state here a weaker version that is suitable for our aim.

Lemma 7.

Let v H 1 ( B 1 ) be a solution of

- div ( A u ) = div G in  𝒟 ( B 1 ) ,

where

G + := 𝟙 H G C 0 , α ( H B 1 ) , G - := 𝟙 H c G C 0 , α ( H c B 1 ) ,

for some α > 0 and A is an elliptic matrix satisfying

ν | z | 2 A i j ( x ) z i z j N | z | 2

and

A + := 𝟙 H A C 0 , α ( H ¯ B 1 ) , A - := 𝟙 H c A C 0 , α ( H ¯ c B 1 ) ,

for some ν , N > 0 . Let us denote

C A = max { A + C 0 , α , A - C 0 , α } , C G = max { G + C 0 , α , G - C 0 , α } .

Then v L loc 2 , n ( B 1 ) (see equation (2.1)). Moreover, there exist two positive constants C = C ( n , ν , N , C A , C G ) and r 0 = r 0 ( n , ν , N , G L , C A , C G ) such that, for any r < r 0 with B r ( x 0 ) B 1 ,

B ρ ( x 0 ) | v | 2 𝑑 x C ( ρ r ) n B r ( x 0 ) | v | 2 𝑑 x + C ρ n for all  ρ < r 4 .

Proof.

Fix x 0 B 1 and let r be such that B r ( x 0 ) B 1 . Let us denote by a + and a - the averages of A in H B r ( x 0 ) and H c B r ( x 0 ) , respectively. In an analogous way we define g + and g - the averages of G in H B r ( x 0 ) and H c B r ( x 0 ) . For x B r ( x 0 ) we define

A ¯ := a + 𝟙 H + a - 𝟙 H c , G ¯ := g + 𝟙 H + g - 𝟙 H c .

Notice that by assumption

(3.10) | A ( x ) - A ¯ ( x ) | C A r α and | G ( x ) - G ¯ ( x ) | C G r α .

Let w be the solution of

{ - div ( A ¯ w ) = div G ¯ in  B r ( x 0 ) , w = v on  B r ( x 0 ) .

The last equation can be rewritten as

(3.11) { - div ( a + w + ) = 0 in  B r ( x 0 ) H , - div ( a - w - ) = 0 in  B r ( x 0 ) H c , w + = w - on  B r ( x 0 ) H , a + w + , e n - a - w - , e n = g + , e n - g - , e n on  B r ( x 0 ) H ,

where w + := w 𝟙 B r ( x 0 ) H , w - := w 𝟙 B r ( x 0 ) H c . Set

D ¯ c w := i = 1 n A ¯ i n i w + G ¯ , e n ,

where A ¯ i n is the ( i , n ) -th entry of the matrix A ¯ . We notice that D ¯ c w has no jumps on the boundary thanks to the transmission condition in (3.11). This allows us to prove that the distributional gradient of D ¯ c w coincides with the point-wise one.

Step 1: Tangential derivatives of w. Let us denote with τ the general direction tangent to the hyperplane H . Since A ¯ and G ¯ are both constant along the tangential directions, the classical difference quotient method gives that τ w W loc 1 , 2 ( B r ( x 0 ) ) and

div ( A ¯ ( τ w ) ) = 0 in  B r ( x 0 ) .

Hence, Caccioppoli’s inequality holds:

(3.12) B ρ ( x ) | ( τ w ) | 2 𝑑 y c ( n , ν , N ) ρ 2 B 2 ρ ( x ) | τ w - ( τ w ) x , 2 ρ | 2 𝑑 y ,

for all balls B 2 ρ ( x ) B r ( x 0 ) and, by De Giorgi’s regularity theorem, τ w is Hölder continuous and there exists γ = γ ( n , ν , N ) > 0 such that if B s ( x ) B r ( x 0 ) ,

(3.13) B ρ ( x ) | τ w - ( τ w ) x , ρ | 2 𝑑 y c ( n , ν , N ) ( ρ s ) n + 2 γ B s ( x ) | τ w - ( τ w ) x , s | 2 𝑑 y ,

for any ρ ( 0 , s 2 ) and

(3.14) max B ρ / 2 ( x ) | τ w | 2 c ( n , ν , N ) ρ n B ρ ( x ) | τ w | 2 𝑑 y .

Step 2: Regularity of D ¯ c w . First of all observe that τ ( D ¯ c w ) = D ¯ c ( τ w ) - G ¯ , e n . This implies by Step 1 that the tangential derivatives of D ¯ c w belong to L loc 2 ( B r ( x 0 ) ) . Furthermore, we can estimate directly by definition of D ¯ c w :

| n ( D ¯ c w ) | c ( n , N ) | τ w | ,

which implies again by Step 1

| D ¯ c w | c ( n , N ) | τ w | .

We can conclude that D ¯ c w W loc 1 , 2 ( B r ( x 0 ) ) . Using Poincaré’s inequality and (3.12), we have

B ρ ( x ) | D ¯ c w - ( D ¯ c w ) x , ρ | 2 𝑑 y c ( n ) ρ 2 B ρ ( x ) | ( D ¯ c w ) | 2 𝑑 y
c ( n , N ) ρ 2 B ρ ( x ) | ( τ w ) | 2 𝑑 y
c ( n , ν , N ) B 2 ρ ( x ) | τ w - ( τ w ) x , 2 ρ | 2 𝑑 y ,

for any B 2 ρ ( x ) B r ( x 0 ) . By (3.13) we infer

B ρ ( x ) | D ¯ c w - ( D ¯ c w ) x , ρ | 2 𝑑 y c ( n , ν , N ) ( ρ r ) n + 2 γ B r / 2 ( x ) | τ w - ( τ w ) x , r 2 | 2 𝑑 y
c ( n , ν , N ) ( ρ r ) n + 2 γ B r ( x 0 ) | τ w | 2 𝑑 y ,

for any x B r / 4 ( x 0 ) and ρ r 4 . Hence by [23, Lemma 4.2] (see also [3, Lemma 7.51]), D ¯ c w is Hölder continuous and by (3.14) we get

(3.15)

max B B r / 4 ( x 0 ) | D ¯ c w | 2 c ( n , ν , N ) B r ( x 0 ) | τ w | 2 𝑑 y + | B r / 4 ( x 0 ) D ¯ c w ( y ) 𝑑 y | 2
c ( n , ν , N ) r n B r ( x 0 ) | w | 2 𝑑 y + 2 G L 2 .

Step 3: Comparison between v and w. Subtracting the equation for w from the equation for v we get

B r ( x 0 ) A ¯ i j ( x ) ( i v - i w ) j φ d x = B r ( x 0 ) ( A ¯ i j ( x ) - A i j ( x ) ) i v j φ d x + B r ( x 0 ) ( G ¯ i - G i ) i φ d x

for any φ W 0 1 , 2 ( B r ( x 0 ) ) . Choosing φ = v - w in the previous equation and using assumption (3.10) we have

ν B r ( x 0 ) | v - w | 2 𝑑 x C A r α B r ( x 0 ) | v | 2 𝑑 y + C G r n + α .

Finally, we can estimate

B ρ ( x 0 ) | v | 2 𝑑 y 2 B ρ ( x 0 ) | w | 2 𝑑 y + 2 B ρ ( x 0 ) | v - w | 2 𝑑 y 2 ω n ρ n sup B r / 4 | w | 2 + 2 B ρ ( x 0 ) | v - w | 2 𝑑 y ,

for any ρ r 4 , and observing that

sup B r / 4 ( x 0 ) | w | 2 = sup B r / 4 ( x 0 ) | τ w | 2 + sup B r / 4 ( x 0 ) | n w | 2
c ( n , ν , N ) sup B r / 4 ( x 0 ) | τ w | 2 + c ( ν ) sup B r / 4 ( x 0 ) | D ¯ c w | 2 + c ( ν , G ) ,

by (3.14), (3.15), the minimality of w and Young’s inequality we gain

B ρ ( x 0 ) | v | 2 𝑑 y c ( n , ν , N ) ( ρ r ) n B r ( x 0 ) | w | 2 𝑑 y + c ( n , ν , G , C A , C G ) [ r α B r ( x 0 ) | v | 2 𝑑 y + r n ]
C ( n , ν , N , G , C A , C G ) { [ ( ρ r ) n + r α ] B r ( x 0 ) | v | 2 𝑑 y + r n } ,

which leads to our aim if we apply Lemma 2. ∎

The next lemma is inspired by [13, Proposition 2.4] and is the main result of this section.

Lemma 8.

Let ( E , u ) be a Λ-minimizer of the functional F defined in (1.1). There exists τ 0 ( 0 , 1 ) such that the following statement is true: for all τ ( 0 , τ 0 ) there exists ε 0 = ε 0 ( τ ) > 0 such that if B r ( x 0 ) Ω with r 1 2 n < τ and one of the following conditions holds:

  1. | E B r ( x 0 ) | < ε 0 | B r ( x 0 ) | ,

  2. | B r ( x 0 ) E | < ε 0 | B r ( x 0 ) | ,

  3. there exists a halfspace H such that | ( Δ E H ) B r ( x 0 ) | | B r ( x 0 ) | < ε 0 ,

then

B τ r ( x 0 ) | u | 2 𝑑 x C 6 [ τ n B r ( x 0 ) | u | 2 𝑑 x + r n ] ,

for some positive constant C 6 = C 6 ( n , ν , N , L , L D , u L 2 ( Ω ) ) .

Proof.

Let us fix B r ( x 0 ) Ω and 0 < τ < 1 . Without loss of generality, we may assume that τ < 1 4 and x 0 = 0 . We start proving (i), being the proof of (ii) similar. Let us define

A i j 0 := a i j ( x 0 , u r / 2 ( x 0 ) ) , B i 0 := a i ( x 0 , u r / 2 ( x 0 ) ) , f 0 := a ( x 0 , u r / 2 ( x 0 ) ) ,

and

F 0 ( z ) := A 0 z , z + B 0 , z + f 0 .

Let us denote by v the solution of the following problem:

min w u + H 0 1 ( B r / 2 ) 0 ( w ; B r / 2 ) ,

where

0 ( w ; B r / 2 ) := B r / 2 F 0 ( w ) 𝑑 x .

Now we use the identity

A 0 ξ , ξ - A 0 η , η = A 0 ( ξ - η ) , ξ - η + 2 A 0 η , ξ - η for all  ξ , η n

in order to deduce that

0 ( u ) - 0 ( v ) = B r / 2 [ A 0 u , u - A 0 v , v ] 𝑑 x + B r / 2 B 0 , u - v 𝑑 x
= B r / 2 A 0 ( u - v ) , u - v 𝑑 x + 2 B r / 2 A 0 v , u - v 𝑑 x + B r / 2 B 0 , u - v 𝑑 x .

By the Euler–Lagrange equation for v we deduce that the sum of the last two integrals in the previous identity is zero being also u = v on B r / 2 . Therefore, using the ellipticity assumption of A 0 we finally achieve that

(3.16) ν B r / 2 | u - v | 2 𝑑 x 0 ( u ) - 0 ( v ) .

Now we prove that u is an ω-minimizer of 0 . We start writing

(3.17)

0 ( u ) = ( E , u ) + [ 0 ( u ) - ( E , u ) ]
( E , v ) + [ 0 ( u ) - ( E , u ) ]
= 0 ( v ) + [ 0 ( u ) - ( E , u ) ] + [ ( E , v ) - 0 ( v ) ] .

Estimate of F 0 ( u ) - F ( E , u ) . We use (1.7) and (3.9) to infer

(3.18)

0 ( u ) - ( E , u ) = B r / 2 ( a i j ( x 0 , u r / 2 ( x 0 ) ) - a i j ( x , u ( x ) ) ) i u j u d x
+ B r / 2 ( a i ( x 0 , u r / 2 ( x 0 ) ) - a i ( x , u ( x ) ) ) i u d x
+ B r / 2 ( a ( x 0 , u r / 2 ( x 0 ) ) - a ( x , u ( x ) ) ) 𝑑 x - B r / 2 E G ( x , u , u ) 𝑑 x
c ( n , L D u L 2 ( Ω ) ) ( r 1 2 B r / 2 | u | 2 𝑑 x + r n + 1 2 ) + C ( N , L ) ( B r / 2 E | u | 2 𝑑 x + r n ) ,

where we denoted L D the greatest Lipschitz constant of the data a i j , b i j , a i , b i , a , b defined in (1.5). Now we use Hölder’s inequality and Lemma 4 to estimate

(3.19) B r / 2 E | u | 2 𝑑 x | E B r | 1 - 1 s | B r | 1 s ( B r / 2 | u | 2 s ) 1 s C 2 1 s ( | E B r | | B r | ) 1 - 1 s B r ( 1 + | u | 2 ) 𝑑 x .

Merging the last estimate in (3.18) we deduce

(3.20)

0 ( u ) - ( E , u ) ( c ( n , L D , u L 2 ( Ω ) ) + C ( N , L ) C 2 1 s ) ( r 1 2 + ε 0 1 - 1 s ) B r | u | 2 𝑑 x
+ ( C 2 1 s + 2 L + c ( n , L D , u L 2 ( Ω ) ) ) r n .

Estimate of F ( E , v ) - F 0 ( v ) . We have

(3.21)

( E , v ) - 0 ( v ) = B r / 2 ( a i j ( x , v ( x ) ) - a i j ( x 0 , u r / 2 ( x 0 ) ) ) i v j v d x
+ B r / 2 ( a i ( x , v ( x ) ) - a i ( x 0 , u r / 2 ( x 0 ) ) ) i v d x
+ B r / 2 ( a ( x , v ( x ) ) - a ( x 0 , u r / 2 ( x 0 ) ) ) 𝑑 x + B r / 2 E G ( x , v , v ) 𝑑 x .

If we choose now z B r / 2 , recalling that u ( z ) = v ( z ) we deduce

| a i j ( x , v ( x ) ) - a i j ( x 0 , u r / 2 ( x 0 ) ) | = | a i j ( x , v ( x ) ) - a i j ( x , v ( z ) ) + a i j ( x , u ( z ) ) - a i j ( x 0 , u r / 2 ( x 0 ) ) |
L D ( | v ( x ) - v ( z ) | + C r 1 2 u L 2 ( Ω ) + r )
L D ( osc ( u , B r / 2 ) + C ( n , ν , N , L ) r + C r 1 2 u L 2 ( Ω ) + r )
C ( n , ν , N , L , L D , u L 2 ( Ω ) ) r 1 2 ,

where we used the fact that osc ( v , B r / 2 ) osc ( u , B r / 2 ) + C ( n , ν , N , L ) r (see [14, Lemma 8.4]). Analogously we can estimate the other differences in (3.21), deducing

( E , v ) - 0 ( v ) C ( n , ν , N , L , L D , u L 2 ( Ω ) ) r 1 2 ( B r / 2 | v | 2 𝑑 x + r n ) + C ( N , L ) ( B r / 2 E | v | 2 𝑑 x + r n ) .

Reasoning in a similar way as in (3.19), we can apply the higher integrability for v given by Lemma 6 and infer that

B r / 2 E | v | 2 𝑑 x C ( n , ν , N , L ) ε 0 1 - 1 s ( B r | u | 2 𝑑 x + r n ) .

Therefore we obtain

(3.22) ( E , v ) - 0 ( v ) C ( n , ν , N , L , L D , u L 2 ( Ω ) ) [ ( r 1 2 + ε 0 1 - 1 s ) B r | u | 2 d x + r n ] .

Finally, collecting (3.16), (3.17), (3.20) and (3.22), if we choose ε 0 such that ε 0 1 - 1 s = τ n , recalling that r 1 2 n < τ , we conclude that

(3.23) B r / 2 | u - v | 2 𝑑 x C [ τ n B r | u | 2 𝑑 x + r n ] ,

for some constant C = C ( n , ν , N , L , L D , u L 2 ( Ω ) ) . On the other hand v is the solution of a uniformly elliptic equation with constant coefficients, so we have

(3.24) B τ r | v | 2 𝑑 x C ( n , ν , N ) τ n B r / 2 | v | 2 𝑑 x C ( n , ν , N , L ) [ τ n B r / 2 | u | 2 𝑑 x + r n ] .

Hence we may estimate, using (3.23) and (3.24),

B τ r | u | 2 𝑑 x 2 B τ r | v - u | 2 𝑑 x + 2 B τ r | v | 2 𝑑 x C [ τ n B r | u | 2 𝑑 x + r n ] ,

for some constant C = C ( n , ν , N , L , L D , u L 2 ( Ω ) ) .

We are left with case (iii). Let H be the half-space from our assumption and let us denote accordingly

A i j 0 ( x ) := a i j ( x , u ( x ) ) + 𝟙 H b i j ( x , u ( x ) ) ,
B i j 0 ( x ) := a i ( x , u ( x ) ) + 𝟙 H b i ( x , u ( x ) ) ,
f 0 ( x ) := a ( x , u ( x ) ) + 𝟙 H b ( x , u ( x ) ) ,
F 0 ( x , z ) := A 0 ( x ) z , z + B 0 ( x ) , z + f 0 ( x ) .

Let us denote by v H the solution of the problem

min w u + H 0 1 ( B r / 2 ) 0 ( w ; B r / 2 ) ,

where

0 ( w ; B r / 2 ) := B r / 2 F 0 ( x , w ) 𝑑 x .

Let us point out that v H solves the Euler–Lagrange equation

(3.25) - 2 div ( A 0 v H ) = div B 0 in  𝒟 ( B r / 2 ) .

Therefore we are in a position to apply Lemma 7 to the function v H . Indeed, from the Hölder continuity of u (see Remark 2) we deduce that the restrictions of A 0 and B 0 onto H B r and B r H respectively are Hölder continuous. We can conclude using also (3.9) that there exist two constants C = C ( n , ν , N , L , L D , u L 2 ( Ω ) ) and τ 0 = τ 0 ( n , ν , N , L , L D , u L 2 ( Ω ) ) such that for τ < τ 0 ,

(3.26) B τ r | v H | 2 𝑑 x C [ τ n B r / 2 | v H | 2 𝑑 x + r n ] .

In addition, using the ellipticity condition of A 0 we can argue as in (3.1) to deduce using also the fact that v H satisfies (3.25),

(3.27) ν B r / 2 | u - v H | 2 𝑑 x 0 ( u ) - 0 ( v H ) .

One more time we can prove that u is an ω-minimizer of 0 . We start as above writing

0 ( u ) = ( E , u ) + [ 0 ( u ) - ( E , u ) ]
( E , v H ) + [ 0 ( u ) - ( E , u ) ]
= 0 ( v H ) + [ 0 ( u ) - ( E , u ) ] + [ ( E , v H ) - 0 ( v H ) ] .

We can estimate the differences 0 ( u ) - ( E , u ) and ( E , v H ) - 0 ( v H ) exactly as before using this time the higher integrability given in Lemma 6. We conclude that

B r / 2 | u - v H | 2 𝑑 x C [ τ n B r | u | 2 𝑑 x + r n ] ,

for some constant C = C ( n , ν , N , L , L D , u L 2 ( Ω ) ) . From the last estimate we can conclude the proof as before using (3.26) and (3.27). ∎

4 Energy density estimates

This section is devoted to proving a lower bound estimate for the functional ( E , u ; B r ( x 0 ) ) . Lemma 8 is the main tool to achieve such result. We shall prove that the energy decays “fast” if the perimeter of E is “small”. In this section we will use a scaling argument.

Scaling.

Let B r ( x 0 ) Ω and let ( E , u ) be a Λ-minimizer of in B r ( x 0 ) . Let us define

E r := E - x 0 r and u r ( y ) := u ( x 0 + r y ) r for all  y B 1 .

By direct computations we deduce that the couple ( E r , u r ) is a Λ r -minimizer in B 1 of the functional

(4.1) r ( E , w ; B 1 ) := r B 1 [ F ( x 0 + r y , r 1 2 w , r - 1 2 w ) + 𝟙 E G ( x 0 + r y , r 1 2 w , r - 1 2 w ) ] 𝑑 y + P ( E , B 1 ) .

Lemma 9.

Let ( E , u ) be a Λ-minimizer of the functional F defined in (1.1). For every τ ( 0 , 1 ) there exists ε 1 = ε 1 ( τ ) > 0 such that if B r ( x 0 ) Ω and P ( E , B r ( x 0 ) ) < ε 1 r n - 1 , then

( E , u ; B τ r ( x 0 ) ) C 7 τ n ( ( E , u ; B r ( x 0 ) ) + r n ) ,

for some positive constant C 7 = C 7 ( n , ν , N , L , L D , Λ , u L 2 ( Ω ) ) independent of τ and r.

Proof.

Let τ ( 0 , 1 ) and B r ( x 0 ) Ω . Without loss of generality, we may assume that τ < 1 2 . We may also assume that x 0 = 0 , and r = 1 by scaling E r = E - x 0 r , u r ( y ) = r - 1 2 u ( x 0 + r y ) for y B 1 , and replacing Λ with Λ r . Thus, we have that ( E r , u r ) is a Λ r -minimizer of r in Ω - x 0 r . For simplicity of notation we can still denote E r by E, u r by u and then we have to prove that there exists ε 1 = ε 1 ( τ ) such that, if P ( E , B 1 ) < ε 1 , then

r ( E , u ; B τ ) C 7 τ n ( r ( E , u ; B 1 ) + r ) .

Note that, since P ( E , B 1 ) is small, by the relative isoperimetric inequality, either | B 1 E | or | B 1 E | is small. Thus Lemma 8 can be applied. Assuming that | B 1 E | is small and using the relative isoperimetric inequality we can deduce that

| B 1 E | c ( n ) P ( E , B 1 ) n n - 1 .

If we choose as a representative of E the set of points of density one, we get by Fubini’s theorem that

| B 1 E | τ 2 τ n - 1 ( B ρ E ) 𝑑 ρ .

Combining these inequalities, we can choose ρ ( τ , 2 τ ) such that

(4.2) n - 1 ( B ρ E ) c ( n ) τ P ( E , B 1 ) n n - 1 c ( n ) ε 1 1 n - 1 τ P ( E , B 1 ) .

Now we set F = E B ρ and observe that

P ( F , B 1 ) P ( E , B 1 B ¯ ρ ) + n - 1 ( B ρ E ) .

If we choose ( F , u ) to test the Λ r -minimality of ( E , u ) we get

r ( E , u ) r ( F , u ) + Λ r | F E |
P ( E , B 1 B ¯ ρ ) + n - 1 ( B ρ E ) + Λ r | B ρ |
+ r B 1 ( F ( x 0 + r y , r 1 2 u ( y ) , r - 1 2 u ( y ) ) + 𝟙 F G ( x 0 + r y , r 1 2 u ( y ) , r - 1 2 u ( y ) ) ) 𝑑 x .

Then getting rid of the common terms, we obtain

P ( E , B ρ ) n - 1 ( B ρ E ) + r B ρ G ( x 0 + r y , r 1 2 u ( y ) , r - 1 2 u ( y ) ) 𝑑 x + Λ r | B ρ | .

Now if we choose ε 1 such that c ( n ) ε 1 1 n - 1 τ n + 1 , we have from (4.2) that

P ( E , B ρ ) τ n P ( E , B 1 ) + r B ρ G ( x 0 + r y , r 1 2 u ( y ) , r - 1 2 u ( y ) ) 𝑑 x + Λ r | B ρ | .

We choose ε 1 satisfying c ( n ) ε 1 n n - 1 ε 0 ( 2 τ ) | B 1 | to obtain, using Lemma 8 and growth conditions (1.6)–(1.7),

r B ρ G ( x 0 + r y , r 1 2 u ( y ) , r - 1 2 u ( y ) ) 𝑑 x C ( N , L ) B ρ ( | u | 2 + r ) 𝑑 x
C ( n , ν , N , L , L D , u L 2 ( Ω ) ) τ n B 1 ( | u | 2 + r ) 𝑑 x .

Finally, we recall that ρ ( τ , 2 τ ) to get

P ( E , B τ ) τ n C ( n , ν , N , L , L D , u L 2 ( Ω ) ) τ n [ B 1 ( | u | 2 + r ) 𝑑 x + P ( E , B 1 ) ] + Λ r | B 2 τ |
C ( n , ν , N , L , L D , Λ , u L 2 ( Ω ) ) τ n [ r ( E , u ; B 1 ) + r ] .

From this estimate the result easily follows applying again Lemma 8. ∎

Theorem 5 (Density lower bound).

Let ( E , u ) be a Λ-minimizer of F in Ω and U Ω an open set. There exists a constant C 8 = C 8 ( n , ν , N , L , L D , Λ , u L 2 ( Ω ) , U ) such that, for every x 0 E and B r ( x 0 ) U , it holds

P ( E , B r ( x 0 ) ) C 8 r n - 1 .

Moreover, H n - 1 ( ( E * E ) Ω ) = 0 .

Proof.

The proof follows from Lemma 8 and Lemma 9 in a standard way, see [13, Proposition 4.4] or [3, Theorem 7.21]. ∎

5 Compactness for sequences of minimizers

In this section we basically follow the path given in [22, Part III]. We start proving a standard compactness result.

Lemma 10 (Compactness).

Let ( E h , u h ) be a sequence of Λ h -minimizers of the functional F in Ω such that sup h F ( E , u h ; Ω ) < and Λ h Λ R + . There exist a (not relabelled) subsequence and a Λ-minimizer of F in Ω, ( E , u ) , such that for every open set U Ω , it holds

E h E in  L 1 ( U ) , u h u in  H 1 ( U ) , P ( E h , U ) P ( E , U ) .

In addition,

(5.1) if  x h E h U and  x h x U , then  x E U ,
(5.2) if  x E U , there exists  x h E h U such that  x h x .

Finally, if we assume also that u h 0 weakly in L loc 2 ( Ω , R n ) and Λ h 0 as h , then E is a local minimizer of the perimeter, that is,

P ( E , B r ( x 0 ) ) P ( F , B r ( x 0 ) ) ,

for every F such that F Δ E B r ( x 0 ) Ω .

Proof.

We start observing that, by the boundedness condition on ( E h , u h ; Ω ) , we may assume that u h weakly converges to u in H 1 ( U ) and strongly in L 2 ( U ) , and 𝟙 E h converges to 𝟙 E in L 1 ( U ) , as h . By lower semicontinuity we are going to prove the Λ-minimality of ( E , u ) . Let us fix B r ( x 0 ) Ω and assume for simplicity of notation that x 0 = 0 . Let ( F , v ) be a test pair such that F Δ E B r and supp ( u - v ) B r . We can handle the perimeter term as in [22], that is, eventually passing to a subsequence and using Fubini’s theorem, we may choose ρ < r such that, once again, F Δ E B ρ and supp ( u - v ) B ρ , and, in addition,

n - 1 ( * F B ρ ) = n - 1 ( * E h B ρ ) = 0 ,

and

(5.3) lim h 0 n - 1 ( B ρ ( Δ E E h ) ) = 0 .

Now we choose a cut-off function ψ C 0 1 ( B r ) such that ψ 1 in B ρ and define

v h := ψ v + ( 1 - ψ ) u h , F h := ( F B ρ ) ( E h B ρ )

to test the minimality of ( E h , u h ) . Thanks to the Λ h -minimality of ( E h , u h ) we have

(5.4)

B r ( F ( x , u h , u h ) + 𝟙 E h G ( x , u h , u h ) ) 𝑑 x + P ( E h , B r )
B r ( F ( x , v h , v h ) + 𝟙 F h G ( x , v h , v h ) ) 𝑑 x + P ( F h , B r ) + Λ h | Δ F h E h |
B r ( F ( x , v h , v h ) + 𝟙 F h G ( x , v h , v h ) ) 𝑑 x + P ( F , B ρ ) + Λ h | Δ F h E h | + P ( E h , B r B ¯ ρ ) + ε h .

The mismatch term ε h = n - 1 ( B ρ ( Δ F ( 1 ) E h ( 1 ) ) ) appears because F is not in general a compact variation of E h . Nevertheless, we have that ε h 0 because of the assumption (5.3) (see also [22, Theorem 21.14]).

Now we use the convexity of F and G with respect to the z variable to deduce

B r ( F ( x , v h , v h ) + 𝟙 F h G ( x , v h , v h ) ) 𝑑 x
B r ( F ( x , v h , ψ v + ( 1 - ψ ) u h ) + 𝟙 F h G ( x , v h , ψ v + ( 1 - ψ ) u h ) ) 𝑑 x
+ B r z F ( x , v h , v h ) , ψ ( v - u h ) 𝑑 x + B r 𝟙 F h z G ( x , v h , v h ) , ψ ( v - u h ) 𝑑 x ,

where the last two terms in the previous estimate tend to zero as h . Indeed, the term ψ ( v - u h ) strongly converges to zero in L 2 , being u = v in B r B ρ and the first part in the scalar product weakly converges in L 2 . Then using again the convexity of F and G with respect to the z variable we obtain, for some infinitesimal σ h ,

(5.5)

B r ( F ( x , v h , v h ) + 𝟙 F h G ( x , v h , v h ) ) 𝑑 x B r ψ ( F ( x , v h , v ) + 𝟙 F h G ( x , v h , v ) ) 𝑑 x
+ B r ( 1 - ψ ) ( F ( x , v h , u h ) + 𝟙 F h G ( x , v h , u h ) ) 𝑑 x + σ h .

Finally, we combine (5.4) and (5.5) and pass to the limit as h + using the lower semicontinuity on the left-hand side. For the right-hand side we observe that 𝟙 E h 𝟙 E and 𝟙 F h 𝟙 F in L 1 ( B r ) and we use also the equi-integrability of { u h } h to conclude

B r ψ ( F ( x , u , u ) + 𝟙 E G ( x , u , u ) ) 𝑑 x + P ( E , B ρ ) B r ψ ( F ( x , v , v ) + 𝟙 F G ( x , v , v ) ) 𝑑 x + P ( F , B ρ ) + Λ | Δ F E | .

Letting ψ 𝟙 B ρ , we finally get

B ρ ( F ( x , u , u ) + 𝟙 E G ( x , u , u ) ) 𝑑 x + P ( E , B ρ )
B ρ ( F ( x , v , v ) + 𝟙 F G ( x , v , v ) ) 𝑑 x + P ( F , B ρ ) + Λ | Δ F E | ,

and this proves the Λ-minimality of ( E , u ) .

To prove the strong convergence of u h to u in L 2 ( B r ) , we start observing that by (5.4) and (5.5) applied using ( E h , u ) to test the Λ-minimality of ( E h , u h ) we get

B r ψ ( F ( x , u h , u h ) + 𝟙 E h G ( x , u h , u h ) ) 𝑑 x B r ψ ( F ( x , u , u ) + 𝟙 E h G ( x , u , u ) ) 𝑑 x + σ h .

Then from the equi-integrability of { u h } h in L 2 ( U ) and recalling that 𝟙 E h 𝟙 E in L 1 ( U ) , we obtain

lim sup h B r ψ ( F ( x , u h , u h ) + 𝟙 E h G ( x , u h , u h ) ) 𝑑 x B r ψ ( F ( x , u , u ) + 𝟙 E G ( x , u , u ) ) 𝑑 x .

The opposite inequality can be obtained by semicontinuity. Thus we get

lim h B r ψ ( F ( x , u h , u h ) + 𝟙 E h G ( x , u h , u h ) ) 𝑑 x = B r ψ ( F ( x , u , u ) + 𝟙 E G ( x , u , u ) ) 𝑑 x .

From the ellipticity condition in (1.6) we infer, for some σ h 0 ,

ν B r ψ | u h - u | 2 𝑑 x B r ψ ( F ( x , u h , u h ) - F ( x , u , u ) ) 𝑑 x
+ B r ψ 𝟙 E ( G ( x , u h , u h ) - G ( x , u , u ) ) 𝑑 x + σ h .

Passing to the limit we obtain

lim h B r ψ | u h - u | 2 𝑑 x = 0 .

Finally, testing the minimality of ( E h , u h ) with the pair ( E , u ) , we also get

lim h P ( E h , B ρ ) = P ( E , B ρ ) .

With a usual argument we can deduce u h u in W 1 , 2 ( U ) and P ( E h , U ) P ( E , U ) for every open set U Ω . The topological information stated in (5.1) and (5.2) follows as in [22, Theorem 21.14] because it does not depend on the presence of the integral bulk part. ∎

6 Decay of the excess

6.1 Excess and height bound

Now we introduce the usual quantities involved in regularity theory. Given x E , a scale r > 0 and a direction ν 𝕊 n - 1 , we define the spherical excess of E with respect to the direction ν as

𝐞 ( x , r , ν ) := 1 r n - 1 E B r ( x ) | ν E ( y ) - ν | 2 2 𝑑 n - 1 ( y ) ,

and the spherical excess of E as

𝐞 ( x , r ) := min ν 𝕊 n - 1 𝐞 ( x , r , ν ) .

In addition, we define the rescaled Dirichlet integral of u in B r ( x 0 ) as

𝒟 ( x , r ) := 1 r n - 1 B r ( x ) | u | 2 𝑑 y .

The following height bound lemma is a standard step in the proof of regularity.

Lemma 11 (Height bound).

Let ( E , u ) be a Λ-minimizer of F in B r ( x 0 ) . There exist two positive constants ε 2 and C 9 , depending on u L 2 ( B r ( x 0 ) ) , such that if x 0 E and

𝐞 ( x , r , ν ) < ε 2 ,

for some ν S n - 1 , then

sup y E B r / 2 ( x 0 ) | ν , y - x 0 | r C 9 𝐞 ( x , r , ν ) 1 2 ( n - 1 ) .

Proof.

The proof of this lemma is almost identical to the one in [22, Theorem 22.8]. Indeed, it follows from the density lower bound (see Theorem 5), the relative isoperimetric inequality and the compactness result proved in the previous section. ∎

Proceeding as in [22] we give the following Lipschitz approximation lemma which is a consequence of the height bound lemma. Its proof follows exactly as in [22, Theorem 23.7].

Theorem 6 (Lipschitz approximation).

Let ( E , u ) be a Λ-minimizer of F in B r ( x 0 ) . There exist two positive constants ε 3 and C 10 , depending on u L 2 ( B r ( x 0 ) ) , such that if x 0 E and

𝐞 ( x 0 , r , e n ) < ε 3 ,

then there exists a Lipschitz function f : R n - 1 R such that

sup x n - 1 | f ( x ) | r C 10 𝐞 ( x 0 , r , e n ) 1 2 ( n - 1 ) , f L 1 ,

and

1 r n - 1 n - 1 ( ( Δ E Γ f ) B r / 2 ( x 0 ) ) C 10 𝐞 ( x 0 , r , e n ) ,

where Γ f is the graph of f. Moreover,

1 r n - 1 B r / 2 n - 1 ( x 0 ) | f | 2 𝑑 x C 10 𝐞 ( x 0 , r , e n ) .

Finally, we shall need the following reverse Poincaré inequality which can be proved exactly as in the case of Λ-minimizers of the perimeter (see [22, Theorem 24.1]).

Theorem 7 (Reverse Poincaré inequality).

Let ( E , u ) be a Λ-minimizer of F in B r ( x 0 ) . There exist two positive constants ε 4 and C 11 such that if x 0 E and

𝐞 ( x 0 , r , ν ) < ε 4 ,

then

𝐞 ( x 0 , r / 2 , ν ) C 11 ( 1 r n + 1 E B r ( x 0 ) | ν , x - x 0 - c | 2 𝑑 n - 1 + 𝒟 ( x 0 , r ) + r ) ,

for every c R .

6.2 Weak Euler–Lagrange equation

The last ingredient to prove the excess improvement is the following Euler–Lagrange equation that we state for Λ r -minimizers of the rescaled functional r defined in (4.1). For the sake of simplicity we will denote as A 1 ( x , s ) the matrix whose entries are a h k ( x , s ) , A 2 ( x , s ) the vector of components a h ( x , s ) , A 3 ( x , s ) = a ( x , s ) and similarly for B i , i = 1 , 2 , 3 . Accordingly, we can write

r ( w , D ) = B 1 [ F r ( x , w , w ) + 𝟙 D G r ( x , w , w ) ] 𝑑 x + P ( D , B 1 )
= B 1 [ ( A 1 r + 𝟙 D B 1 r ) w , w + r A 2 r + 𝟙 D B 2 r , w + r ( A 3 r + 𝟙 D B 3 r ) ] 𝑑 x + P ( D , B 1 ) ,

where r > 0 , x 0 Ω , A i r := A i ( x 0 + r y , r w ) , B i r := B i ( x 0 + r y , r w ) , for i = 1 , 2 , 3 . The argument used to prove the next result is similar to the one in [3, Theorem 7.35].

Theorem 8 (Weak Euler–Lagrange equation).

Let ( E , u ) be a Λ r -minimizer of F r in B 1 . For every vector field X C 0 1 ( B 1 , R n ) and for some constant C 12 = C 12 ( N , L D , sup | X | , sup | X | ) > 0 it holds

(6.1) E div τ X d n - 1 C 12 B 1 ( | u | 2 + r ) 𝑑 x + Λ r E | X | 𝑑 n - 1 ,

where div τ denotes the tangential divergence on E , i.e.,

div τ X = div X - ν E , X ν E .

Proof.

Let us fix X C 0 1 ( B 1 , n ) . We set Φ t ( x ) = x + t X ( x ) , E t = Φ t ( E ) and u t = u Φ t - 1 , for any t > 0 . From the Λ r -minimality it follows that

(6.2)

[ P ( E t , B 1 ) - P ( E , B 1 ) ] + Λ r | Δ E t E | + B 1 [ F r ( y , u t , u t ) + 𝟙 E t ( y ) G r ( y , u t , u t ) ] 𝑑 y
- B 1 [ F r ( x , u , u ) + 𝟙 E ( x ) G r ( x , u , u ) ] 𝑑 x 0 .

In order to obtain (6.1), we will divide by t and pass to the upper limit as t 0 + . Let us study these terms separately. The first variation of the area gives

(6.3) lim t 0 + 1 t [ P ( E t , B 1 ) - P ( E , B 1 ) ] = E div τ X d n - 1 .

We can deal with the second term observing that

(6.4) lim t 0 + | Δ E t E | t E | X , ν E | 𝑑 n - 1

(see for instance [16, Theorem 3.2]). In the first bulk term we make the change of variables y = Φ t ( x ) with x B 1 and t > 0 , taking into account that

Φ t - 1 ( Φ t ( x ) ) = I - t X ( x ) + o ( t ) , J Φ t ( x ) = 1 + t div X ( x ) + o ( t ) .

Thus we gain

B 1 [ F r ( y , u t , u t ) + 𝟙 E t ( y ) G r ( y , u t , u t ) ] 𝑑 y
= B 1 [ F r ( Φ t ( x ) , u , u ) + 𝟙 E ( x ) G r ( Φ t ( x ) , u , u ) ] ( 1 + t div X ) 𝑑 x
- t B 1 [ 2 C 1 u X , u + r C 2 , u X ] 𝑑 x + o ( t ) ,

where we set

C i := A ~ i r + 𝟙 E B ~ i r = A i r ( Φ t ( x ) , u ) + 𝟙 E ( x ) B i r ( Φ t ( x ) , u ) ,

for i = 1 , 2 , 3 . By simple calculations we obtain

B 1 [ F r ( y , u t , u t ) + 𝟙 E t ( y ) G r ( y , u t , u t ) ] 𝑑 y - B 1 [ F r ( x , u , u ) + 𝟙 E ( x ) G r ( x , u , u ) ] 𝑑 x
= B 1 { F r ( Φ t ( x ) , u , u ) + 𝟙 E ( x ) G r ( Φ t ( x ) , u , u ) - [ F r ( x , u , u ) + 𝟙 E ( x ) G r ( x , u , u ) ] } 𝑑 x
+ t [ B 1 [ F r ( Φ t ( x ) , u , u ) + 𝟙 E ( x ) G r ( Φ t ( x ) , u , u ) ] div X d x
- B 1 [ 2 C 1 u X , u + r C 2 , u X ] d x ] + o ( t ) .

Let us estimate the first of the three terms. By Lipschitz continuity and Young’s inequality we get

B 1 { F r ( Φ t ( x ) , u , u ) + 𝟙 E ( x ) G r ( Φ t ( x ) , u , u ) - [ F r ( x , u , u ) + 𝟙 E ( x ) G r ( x , u , u ) ] } d x
c ( L D ) t B 1 | X | [ | u | 2 + r | u | + r ] 𝑑 x c ( L D ) t B 1 | X | [ | u | 2 + r ] 𝑑 x .

Finally, dividing by t and passing to the upper limit as t 0 + we infer

(6.5)

lim sup t 0 + 1 t [ B 1 [ F r ( y , u t , u t ) + 𝟙 E t ( y ) G r ( y , u t , u t ) ] 𝑑 y - B 1 [ F r ( x , u , u ) + 𝟙 E G r ( x , u , u ) ] 𝑑 x ]
c ( L D ) B 1 | X | [ | u | 2 + r ] 𝑑 x + B 1 [ F r ( x , u , u ) + 𝟙 E G r ( x , u , u ) ] div X d x
- B 1 [ 2 ( A 1 r + 𝟙 E B 1 r ) u X , u + r ( A 2 r + 𝟙 E B 2 r ) , u X ] 𝑑 x .

Passing to the upper limit as t 0 + in (6.2) and putting (6.3), (6.4), (6.5) together we get

E div τ X d n - 1 c ( L D ) B 1 | X | [ | u | 2 + r ] 𝑑 x + | B 1 [ F r ( x , u , u ) + 𝟙 E G r ( x , u , u ) ] div X d x |
+ B 1 | 2 ( A 1 r + 𝟙 E B 1 r ) u X , u + r ( A 2 r + 𝟙 E B 2 r ) , u X | 𝑑 x + Λ r E | X | 𝑑 n - 1
C B 1 ( | u | 2 + r ) 𝑑 x + Λ r E | X | 𝑑 n - 1 ,

where C = C ( N , L D , sup | X | , sup | X | ) . ∎

7 Excess improvement

Theorem 9 (Excess improvement).

For every τ ( 0 , 1 2 ) and M > 0 there exists a constant ε 5 = ε 5 ( τ , M ) ( 0 , 1 ) such that if ( E , u ) is a Λ-minimizer of F in B r ( x 0 ) with x 0 E and

(7.1) 𝐞 ( x 0 , r ) ε 5 , 𝒟 ( x 0 , r ) + r M 𝐞 ( x 0 , r ) ,

then there exists a positive constant C 13 , depending on u L 2 ( B r ( x 0 ) ) , such that

𝐞 ( x 0 , τ r ) C 13 ( τ 2 𝐞 ( x 0 , r ) + 𝒟 ( x 0 , 2 τ r ) + τ r ) .

Proof.

Without loss of generality we may assume that τ < 1 8 . Let us rescale and assume by contradiction that there exist an infinitesimal sequence { ε h } h + , a sequence { r h } h + and a sequence { ( E h , u h ) } h of Λ r h -minimizers of r h in B 1 , with equibounded energies, such that, denoting by 𝐞 h the excess of E h and by 𝒟 h the rescaled Dirichlet integral of u h , we have

𝐞 h ( 0 , 1 ) = ε h , 𝒟 h ( 0 , 1 ) + r h M ε h , 𝐞 h ( 0 , τ ) > C 13 ( τ 2 𝐞 ( 0 , 1 ) + 𝒟 ( 0 , 2 τ ) + τ r h ) ,

with some positive constant C 13 to be chosen. Up to rotating each E h we may also assume that, for all h ,

𝐞 h ( 0 , 1 ) = 1 2 E h B 1 | ν E h - e n | 2 𝑑 n - 1 .

Step 1. Thanks to the Lipschitz approximation theorem, for h sufficiently large, there exists a 1-Lipschitz function f h : n - 1 such that

(7.2) sup n - 1 | f h | C 10 ε h 1 2 ( n - 1 ) , n - 1 ( ( Δ E h Γ f h ) B 1 / 2 ) C 10 ε h , B 1 / 2 n - 1 | f h | 2 𝑑 x C 10 ε h .

We define

g h ( x ) := f h ( x ) - a h ε h , where  a h = B 1 / 2 n - 1 f h 𝑑 x ,

and we assume, up to a subsequence, that { g h } h converges weakly in H 1 ( B 1 / 2 n - 1 ) and strongly in L 2 ( B 1 / 2 n - 1 ) to a function g.

We prove that g is harmonic in B 1 / 2 n - 1 . It is enough show that

(7.3) lim h 1 ε h B 1 / 2 n - 1 f h , ϕ 1 + | f h | 2 𝑑 x = 0 ,

for all ϕ C 0 1 ( B 1 / 2 n - 1 ) ; indeed, if ϕ C 0 1 ( B 1 / 2 n - 1 ) , by weak convergence we have

B 1 / 2 n - 1 g , ϕ 𝑑 x = lim h 1 ε h B 1 / 2 n - 1 f h , ϕ 𝑑 x
= lim h 1 ε h { B 1 / 2 n - 1 f h , ϕ 1 + | f h | 2 𝑑 x + B 1 / 2 n - 1 [ f h , ϕ - f h , ϕ 1 + | f h | 2 ] 𝑑 x } .

Using the Lipschitz continuity of f h and the third inequality in (7.2), we infer that the second term in the previous equality is infinitesimal:

lim sup h 1 ε h | B 1 / 2 n - 1 [ f h , ϕ - f h , ϕ 1 + | f h | 2 ] 𝑑 x | lim sup h 1 ε h B 1 / 2 n - 1 | f h | | ϕ | 1 + | f h | 2 - 1 1 + | f h | 2 𝑑 x
lim sup h 1 ε h B 1 / 2 n - 1 | ϕ | | f h | 2 𝑑 x
lim h C 10 ϕ ε h = 0 .

Therefore, we can prove (7.3). We fix δ > 0 so that spt ϕ × [ - 2 δ , 2 δ ] B 1 / 2 , choose a cut-off function ψ : [ 0 , 1 ] with spt ψ ( - 2 δ , 2 δ ) , ψ = 1 in ( - δ , δ ) and apply to E h the weak Euler–Lagrange equation with X = ϕ ψ e n . By the height bound, for h sufficiently large it holds that E h B 1 / 2 B 1 / 2 n - 1 × ( - δ , δ ) . Plugging X in the weak Euler–Lagrange equation and using the assumption in (7.1), we have

- E h B 1 / 2 ν E h , e n ϕ , ν E h 𝑑 n - 1 c ( N , L D , ϕ , ψ ) B 1 / 2 ( | u h | 2 + r h ) 𝑑 x + Λ r h E h B 1 / 2 | ϕ ψ | 𝑑 n - 1
c ( n , N , Λ , L D , M , ϕ , ψ ) ε h .

Therefore, if we replace ϕ by - ϕ , we infer

(7.4) lim h 1 ε h E h B 1 / 2 ν E h , e n ϕ , ν E h 𝑑 n - 1 = 0 .

Decomposing E h B 1 / 2 = ( [ Γ f h ( E h Γ f h ) ] ( Γ f h E h ) ) B 1 / 2 , we deduce

- 1 ε h E h B 1 / 2 ν E h , e n ϕ , ν E h d n - 1 = 1 ε h [ - Γ f h B 1 / 2 ν E h , e n ϕ , ν E h d n - 1
- ( E h Γ f h ) B 1 / 2 ν E h , e n ϕ , ν E h 𝑑 n - 1
+ ( Γ f h E h ) B 1 / 2 ν E h , e n ϕ , ν E h d n - 1 ] .

Since by the second inequality in (7.2) we have

| 1 ε h ( E h Γ f h ) B 1 / 2 ν E h , e n ϕ , ν E h 𝑑 n - 1 | C 10 ε h sup n - 1 | ϕ | ,

| 1 ε h ( Γ f h E h ) B 1 / 2 ν E h , e n ϕ , ν E h 𝑑 n - 1 | C 10 ε h sup n - 1 | ϕ | ,

by (7.4) and the area formula, we infer

0 = lim h - 1 ε h Γ f h B 1 / 2 ν E h , e n ϕ , ν E h 𝑑 n - 1 = lim h 1 ε h B 1 / 2 n - 1 f h , ϕ 1 + | f h | 2 𝑑 x .

This proves that g is harmonic.

Step 2. The proof of this step now follows exactly as in [13] using the height bound lemma and the reverse Poincaré inequality. We give it here to be thorough. By the mean value property of harmonic functions, [22, Lemma 25.1], Jensen’s inequality, semicontinuity and the third inequality in (7.2) we deduce that

lim h 1 ε h B 2 τ n - 1 | f h ( x ) - ( f h ) 2 τ - ( f h ) 2 τ , x | 2 𝑑 x = B 2 τ n - 1 | g ( x ) - ( g ) 2 τ - ( g ) 2 τ , x | 2 𝑑 x
= B 2 τ n - 1 | g ( x ) - g ( 0 ) - g ( 0 ) , x | 2 𝑑 x
c ( n ) τ n - 1 sup x B 2 τ n - 1 | g ( x ) - g ( 0 ) - g ( 0 ) , x | 2
c ( n ) τ n + 3 B 1 / 2 n - 1 | g | 2 𝑑 x
c ( n ) τ n + 3 lim inf h B 1 / 2 n - 1 | g h | 2 𝑑 x
C ~ ( n , C 10 ) τ n + 3 .

On one hand, using the area formula, the mean value property, the previous inequality and setting

c h := ( f h ) 2 τ 1 + | ( f h ) 2 τ | 2 , ν h := ( - ( f h ) 2 τ , 1 ) 1 + | ( f h ) 2 τ | 2 ,

we have

lim sup h 1 ε h E h Γ f h B 2 τ | ν h , x - c h | 2 𝑑 n - 1
= lim sup h 1 ε h E h Γ f h B 2 τ | - ( f h ) 2 τ , x + f h ( x ) - ( f h ) 2 τ | 1 + | ( f h ) 2 τ | 2 2 1 + | f h ( x ) | 2 𝑑 x
lim h 1 ε h B 2 τ n - 1 | f h ( x ) - ( f h ) 2 τ - ( f h ) 2 τ , x | 2 𝑑 x C ~ ( n , C 10 ) τ n + 3 .

On the other hand, arguing as in Step 1, we immediately get from the height bound and the first two inequalities in (7.2) that

lim h 1 ε h ( E h Γ f h ) B 2 τ | ν h , x - c h | 2 𝑑 n - 1 = 0 .

Hence we conclude that

(7.5) lim sup h 1 ε h E h B 2 τ | ν h , x - c h | 2 𝑑 n - 1 C ~ ( n , C 10 ) τ n + 3 .

We claim that the sequence { 𝐞 h ( 0 , 2 τ , ν h ) } h is infinitesimal; indeed, by the definition of excess, Jensen’s inequality and the third inequality in (7.2) we have

lim sup h E h B 2 τ | ν E h - ν h | 2 𝑑 n - 1 lim sup h [ 2 E h B 2 τ | ν E h - e n | 2 𝑑 n - 1 + 2 | e n - ν h | 2 n - 1 ( E h B 2 τ ) ]
lim sup h [ 4 ε h + 2 n - 1 ( B 2 τ ) | ( ( f h ) 2 τ , 1 + | ( f h ) 2 τ | 2 - 1 ) | 2 1 + | ( f h ) 2 τ | 2 ]
lim sup h [ 4 ε h + 4 n - 1 ( B 2 τ ) | ( f h ) 2 τ | 2 ]
lim sup h [ 4 ε h + 4 B 1 / 2 n - 1 | f h | 2 𝑑 x ] lim h [ 4 ε h + 4 C 10 ε h ] = 0 .

Therefore, applying the reverse Poincaré inequality and (7.5), we have for h large that

𝐞 h ( 0 , τ ) 𝐞 h ( 0 , τ , ν h ) C 11 ( C ~ τ 2 𝐞 h ( 0 , 1 ) + 𝒟 ( 0 , 2 τ ) + 2 τ r h ) ,

which is a contradiction if we choose C 13 > C 11 max { C ~ , 2 } . ∎

8 Proof of the main theorem

The proof works exactly as in [13]. We give here some details to emphasize the dependence of the constant ε appearing in the statement of Theorem 1 from the structural data of the functional. The proof is divided into four steps.

Step 1.

We show that for every τ ( 0 , 1 ) there exists ε 6 = ε 6 ( τ ) > 0 such that if 𝐞 ( x , r ) ε 6 , then

𝒟 ( x , τ r ) C 6 τ 𝒟 ( x , r ) ,

where C 6 is from Lemma 8. Assume by contradiction that for some τ ( 0 , 1 ) there exist two positive sequences ( ε h ) h and ( r h ) h and a sequence ( E h , u h ) of Λ r h -minimizers of r h in B 1 with equibounded energies such that, denoting by 𝐞 h the excess of E h and by 𝒟 h the rescaled Dirichlet integral of u h , we have that 0 E h ,

(8.1) 𝐞 h ( 0 , 1 ) = ε h 0 and 𝒟 h ( 0 , τ ) > C 6 τ 𝒟 h ( 0 , 1 ) .

Thanks to the energy upper bound (Theorem 3) and the compactness lemma (Lemma 10) we may assume that E h E in L 1 ( B 1 ) and 0 E . Since, by lower semicontinuity, the excess of E at 0 is null, it follows that E is a half-space in B 1 , say H. In particular, for h large it holds

| ( Δ E h H ) B 1 | < ε 0 ( τ ) | B 1 | ,

where ε 0 is from Lemma 8, which gives a contradiction with the inequality (8.1).

Step 2.

Let U Ω be an open set. Prove that for every τ ( 0 , 1 ) there exist two positive constants ε 7 = ε 7 ( τ , U ) and C 14 such that if x 0 E , B r ( x 0 ) U and 𝐞 ( x 0 , r ) + 𝒟 ( x 0 , r ) + r < ε 7 , then

(8.2) 𝐞 ( x 0 , τ r ) + 𝒟 ( x 0 , τ r ) + τ r C 14 τ ( 𝐞 ( x 0 , r ) + 𝒟 ( x 0 , r ) + r ) .

Fix τ ( 0 , 1 ) and assume without loss of generality that τ < 1 2 . We can distinguish two cases.

Case 1: D ( x 0 , r ) + r τ - n e ( x 0 , r ) . If 𝐞 ( x 0 , r ) < min { ε 5 ( τ , τ - n ) , ε 6 ( 2 τ ) } , it follows from Theorem 9 and Step 1 that

𝐞 ( x 0 , τ r ) C 13 ( τ 2 𝐞 ( x 0 , r ) + 𝒟 ( x 0 , 2 τ r ) + τ r ) C 13 τ ( 𝐞 ( x 0 , r ) + 2 C 6 𝒟 ( x 0 , r ) + r ) .

Case 2: e ( x 0 , r ) τ n ( D ( x 0 , r ) + r ) . By the property of the excess at different scales, we infer

𝐞 ( x 0 , τ r ) τ 1 - n 𝐞 ( x 0 , r ) τ ( 𝒟 ( x 0 , r ) + r ) .

We conclude that choosing ε 7 = min { ε 5 ( τ , τ - n ) , ε 6 ( 2 τ ) , ε 6 ( τ ) } , inequality (8.2) is verified.

Step 3.

Fix σ ( 0 , 1 2 ) and choose τ 0 ( 0 , 1 ) such that C 14 τ 0 τ 0 2 σ . Let U Ω be an open set. We define

Γ U := { x E U : 𝐞 ( x , r ) + 𝒟 ( x , r ) + r < ε 7 ( τ 0 , U ) ,  for some  r > 0  such that  B r ( x ) U } .

Note that Γ U is relatively open in E . We show that Γ U is a C 1 , σ -hypersurface. Indeed, inequality (8.2) implies via standard iteration argument that if x 0 Γ U , there exist r 0 > 0 and a neighborhood V of x 0 such that for every x E V it holds

𝐞 ( x , τ 0 k r 0 ) + 𝒟 ( x , τ 0 k r 0 ) + τ 0 k r 0 τ 0 2 σ k , for  k 0 .

In particular, 𝐞 ( x , τ 0 k r 0 ) τ 0 2 σ k and, arguing as in [13], we obtain that for every x E V and 0 < s < t < r 0 it holds

| ( ν E ) s ( x ) - ( ν E ) t ( x ) | c t σ ,

for some constant c = c ( n , τ 0 , r 0 ) , where

( ν E ) t ( x ) = E B t ( x ) ν E 𝑑 n - 1 .

The previous estimate first implies that Γ U is C 1 . By a standard argument we then deduce again from the same estimate that Γ U is a C 1 , σ -hypersurface. Finally, we define Γ := i ( Γ U i ) , where ( U i ) i is an increasing sequence of open sets such that U i Ω and Ω = i U i .

Step 4.

Finally, we are in a position to prove that there exists ε > 0 such that

n - 1 - ε ( E Γ ) = 0 .

Being the argument rather standard, Setting Σ = { x E Γ : lim r 0 𝒟 ( x , r ) = 0 } , by Lemma 4 we have u L loc 2 s ( Ω ) for some s = s ( n , ν , N , L ) > 1 and we have

dim ( { x Ω : lim sup r 0 𝒟 ( x , r ) > 0 } ) n - s .

The conclusion follows as in [13] (see also [6] and [7]) showing that Σ = if n 7 and dim ( Σ ) n - 8 if n 8 .


Communicated by Irene Fonseca


Acknowledgements

The authors wish to thank the reviewers for the constructive comments and remarks. The authors are members of the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM).

References

[1] H. W. Alt and L. A. Caffarelli, Existence and regularity for a minimum problem with free boundary, J. Reine Angew. Math. 325 (1981), 105–144. 10.1515/crll.1981.325.105Search in Google Scholar

[2] L. Ambrosio and G. Buttazzo, An optimal design problem with perimeter penalization, Calc. Var. Partial Differential Equations 1 (1993), no. 1, 55–69. 10.1007/BF02163264Search in Google Scholar

[3] L. Ambrosio, N. Fusco and D. Pallara, Functions of Bounded Variation and Free Discontinuity Problems, Oxford Math. Monogr., Oxford University, New York, 2000. 10.1093/oso/9780198502456.001.0001Search in Google Scholar

[4] M. Carozza, I. Fonseca and A. Passarelli di Napoli, Regularity results for an optimal design problem with a volume constraint, ESAIM Control Optim. Calc. Var. 20 (2014), no. 2, 460–487. 10.1051/cocv/2013071Search in Google Scholar

[5] M. Carozza, I. Fonseca and A. Passarelli di Napoli, Regularity results for an optimal design problem with quasiconvex bulk energies, Calc. Var. Partial Differential Equations 57 (2018), no. 2, Paper No. 68. 10.1007/s00526-018-1343-9Search in Google Scholar

[6] C. De Lellis, M. Focardi and B. Ruffini, A note on the Hausdorff dimension of the singular set for minimizers of the Mumford–Shah energy, Adv. Calc. Var. 7 (2014), no. 4, 539–545. 10.1515/acv-2013-0107Search in Google Scholar

[7] G. De Philippis and A. Figalli, Higher integrability for minimizers of the Mumford–Shah functional, Arch. Ration. Mech. Anal. 213 (2014), no. 2, 491–502. 10.1007/s00205-014-0729-xSearch in Google Scholar

[8] G. De Philippis and A. Figalli, A note on the dimension of the singular set in free interface problems, Differential Integral Equations 28 (2015), no. 5–6, 523–536. 10.57262/die/1427744099Search in Google Scholar

[9] L. Esposito, Density lower bound estimate for local minimizer of free interface problem with volume constraint, Ric. Mat. 68 (2019), no. 2, 359–373. 10.1007/s11587-018-0407-7Search in Google Scholar

[10] L. Esposito and N. Fusco, A remark on a free interface problem with volume constraint, J. Convex Anal. 18 (2011), no. 2, 417–426. Search in Google Scholar

[11] I. Fonseca and N. Fusco, Regularity results for anisotropic image segmentation models, Ann. Sc. Norm. Super. Pisa Cl. Sci. (4) 24 (1997), no. 3, 463–499. Search in Google Scholar

[12] I. Fonseca, N. Fusco, G. Leoni and M. Morini, Equilibrium configurations of epitaxially strained crystalline films: Existence and regularity results, Arch. Ration. Mech. Anal. 186 (2007), no. 3, 477–537. 10.1007/s00205-007-0082-4Search in Google Scholar

[13] N. Fusco and V. Julin, On the regularity of critical and minimal sets of a free interface problem, Interfaces Free Bound. 17 (2015), no. 1, 117–142. 10.4171/IFB/336Search in Google Scholar

[14] E. Giusti, Direct Methods in the Calculus of Variations, World Scientific, River Edge, 2003. 10.1142/5002Search in Google Scholar

[15] M. E. Gurtin, On phase transitions with bulk, interfacial, and boundary energy, Arch. Ration. Mech. Anal. 96 (1986), no. 3, 243–264. 10.1007/BF00251908Search in Google Scholar

[16] V. Julin and G. Pisante, Minimality via second variation for microphase separation of diblock copolymer melts, J. Reine Angew. Math. 729 (2017), 81–117. 10.1515/crelle-2014-0117Search in Google Scholar

[17] L. Lamberti, A regularity result for minimal configurations of a free interface problem, Boll. Unione Mat. Ital. 14 (2021), no. 3, 521–539. 10.1007/s40574-021-00285-6Search in Google Scholar

[18] C. J. Larsen, Regularity of components in optimal design problems with perimeter penalization, Calc. Var. Partial Differential Equations 16 (2003), no. 1, 17–29. 10.1007/s005260100139Search in Google Scholar

[19] H. Li, T. Halsey and A. Lobkovsky, Singular shape of a fluid drop in an electric or magnetic field, Europhys. Lett. 27 (1994), 575–580. 10.1209/0295-5075/27/8/004Search in Google Scholar

[20] F.-H. Lin, Variational problems with free interfaces, Calc. Var. Partial Differential Equations 1 (1993), no. 2, 149–168. 10.1007/BF01191615Search in Google Scholar

[21] F. H. Lin and R. V. Kohn, Partial regularity for optimal design problems involving both bulk and surface energies, Chinese Ann. Math. Ser. B 20 (1999), no. 2, 137–158. 10.1142/S0252959999000175Search in Google Scholar

[22] F. Maggi, Sets of Finite Perimeter and Geometric Variational Problems. An Introduction to Geometric Measure Theory, Cambridge Stud. Adv. Math. 135, Cambridge University, Cambridge, 2012. 10.1017/CBO9781139108133Search in Google Scholar

[23] E. Mukoseeva and G. Vescovo, Minimality of the ball for a model of charged liquid droplets, preprint (2019), https://arxiv.org/abs/1912.07092. Search in Google Scholar

[24] G. I. Taylor, Disintegration of water drops in an electric field, Proc. R. Soc. Lond. A 280 (1964), 383–397. 10.1098/rspa.1964.0151Search in Google Scholar

Received: 2021-10-03
Revised: 2022-07-08
Accepted: 2022-09-05
Published Online: 2022-10-25
Published in Print: 2023-10-01

© 2022 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 23.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/acv-2021-0080/html
Scroll to top button