Skip to content
BY 4.0 license Open Access Published by De Gruyter February 7, 2020

The ε - εβ Property in the Isoperimetric Problem with Double Density, and the Regularity of Isoperimetric Sets

  • Aldo Pratelli EMAIL logo and Giorgio Saracco ORCID logo

Abstract

We prove the validity of the ε-εβ property in the isoperimetric problem with double density, generalising the known properties for the case of single density. As a consequence, we derive regularity for isoperimetric sets.

MSC 2010: 49Q10; 49Q20; 58B20

1 Introduction

For any N2, we consider the isoperimetric problem in N with double density. More precisely, two lower semi-continuous (l.s.c.) and locally summable functions f:N(0,+) and h:N×𝕊N-1(0,+) are given, and we measure the volume and perimeter of any Borel set EN according to the formulae

| E | := E f ( x ) 𝑑 x , P ( E ) := * E h ( x , ν E ( x ) ) 𝑑 N - 1 ( x ) ,

where for every set of locally finite perimeter E we denote by *E its reduced boundary and by νE(x)𝕊N-1 the outer normal at x*E, while P(E)=+ whenever E is not a set of locally finite perimeter (see Section 1.2 for the basic definitions of the theory of sets of finite perimeter). We will refer to the “single density” case when h(x,ν)=f(x) for every xN, and ν𝕊N-1. Whenever using the standard N-dimensional Lebesgue measure (i.e. f1) or Euclidean perimeter (i.e. h1), we shall use the subscript Eucl, i.e. ||Eucl and PEucl().

The isoperimetric problem with single density is a wide generalisation of the classical Euclidean isoperimetric problem, and it has been deeply studied in the last decades, we refer the interested reader to [8, 11, 10, 16, 22, 25, 28, 29, 32] and the references therein. The case of double density is yet a further important generalisation, since many of the possible applications correspond to two different densities. The simplest example is given by Riemannian manifolds, which locally behave as N with double density, being f the norm of the Riemannian metric, and h its derivative. In general, in view of applications it is crucial not only that f and h may differ, but also that h may depend both on the point (from now on, the spatial variable), and on the direction of the tangent space at that point (from now on the angular variable).

The isoperimetric problem consists in finding, if they exist, the sets E of minimal perimeter among those of fixed volume. The three main questions one is interested in are existence, boundedness and regularity of isoperimetric sets. Existence in the single density case has been studied in several papers, some of which are those quoted above; for the case of double density it has been studied either for some rather specific choices of the weights (radial [2, 12, 18, 17], monomial [1, 4, 5, 3], Gauss-like [9, 25] or in some Carnot groups [19, 21]), or under very general conditions on f and h, see [20, 30, 33]. Concerning boundedness and regularity, a quite wide answer has been given in the paper [14] for the single density case. The purpose of the present paper is to generalise the results of that paper to the case of double density.

A fundamental tool to study the isoperimetric problem is the classical “ε-ε” regularity property of Almgren, which basically says that one can locally modify a set E by changing its volume by a small (positive or negative) quantity ε, while increasing the perimeter by at most a quantity C|ε|. In the case of single density, the ε-ε property is true for any set of locally finite perimeter as soon as the density is at least locally Lipschitz, but otherwise it is in general false. This is the main reason why many regularity results for isoperimetric sets require a Lipschitz regularity hypothesis on the density.

To extend the study to the case of less regular densities, it is convenient to weaken the ε-ε property to the so-called ε-εβ property, introduced in [14], that we immediately recall.

Definition 1 (The ε-εβ Property).

Let E be a set of locally finite perimeter and β[0,1]. We say that E possesses the ε-εβ property (relative to the densities f and h) if for any ball B such that N-1(B*E)>0 there exist constants C>0 and ε¯>0 such that for all |ε|<ε¯, there exists a set F such that

(1.1) F E B , | F | - | E | = ε , P ( F ) - P ( E ) C | ε | β .

In the case of single density, in [14] it was shown that whenever f is Hölder continuous with some exponent 0α1, then the ε-εβ property holds for every set of locally finite perimeter for some β=β(N,α). Moreover, if f is locally bounded and continuous, the ε-εβ property holds with β=N-1N and with any positive constant C. This means that for any ball B the constant C>0 in Definition 1 can be taken arbitrarily small, up to choosing ε¯ small enough. As a consequence of the ε-εβ property, boundedness and C1,η regularity of isoperimetric sets for some η=η(N,α) was obtained. We shall show that the results of [14] can be extended to the general case of a double density. More precisely, we have the following three results.

Theorem A (The ε-εβ Property).

Assume that f and h are locally bounded, that h is locally α-Hölder in the spatial variable for some α[0,1], and that ERN is a set of locally finite perimeter. Then E possesses the ε-εβ property, where β is given by

(1.2) β = β ( N , α ) = α + ( N - 1 ) ( 1 - α ) α + N ( 1 - α ) .

If α=0 (in which case locally α-Hölder precisely means locally bounded) and h is continuous in the spatial variable, then E possesses the ε-εN-1N property for all constants C>0 (that is, given any ball B, then the constant C of Definition 1 can be taken arbitrarily small, up to choosing ε¯ small enough).

Notice that the ε-εβ property only requires the regularity of h in the spatial variable, while no regularity of h in the angular variable or of f is required — except that both f and h have to be l.s.c. and locally L1, which is required in general to set the problem.

Theorem B (Boundedness).

Assume that there exists a constant M>0 such that

(1.3) 1 M f ( x ) M , 1 M h ( x , ν ) M for all x N , ν 𝕊 N - 1 ,

and that ERN is an isoperimetric set for which the ε-εN-1N property holds with an arbitrarily small constant C. Then E is bounded.

Notice that this result, paired with Theorem A, ensures the boundedness of isoperimetric sets in a wide generality, i.e. whenever f and h are bounded and are away from 0 – that is, (1.3) holds – and h is continuous in the spatial variable.

Theorem C (Regularity of isoperimetric sets).

Assume that f and h are locally bounded. Then any isoperimetric set is porous (see Definition 1), and its reduced boundary coincides HN-1-a.e. with its topological boundary. In addition, if h=h(x) is locally α-Hölder for some α(0,1], then *E is C1,η, where

(1.4) η = η ( N , α ) = α 2 N ( 1 - α ) + 2 α .

This regularity result is not sharp. In particular, in the 2-dimensional case a higher regularity is shown in [13] for the case of single density. Such an improved regularity in the case of a double density and as well in the case h=h(x,ν) will be addressed in the forthcoming paper [31].

Notice that, since f and h are l.s.c. and positive, they are always locally away from zero. Thus, in Theorems A and C, f and h are locally bounded and locally away from zero, while in Theorem B these requests are made globally. These assumptions are essentially sharp. They can be slightly relaxed to the case of “essentially bounded” or “essentially α-Hölder” functions as defined in [14, Definitions 1.6 and 1.7]. Loosely speaking, this relaxation allows the densities to take the values 0 and + in a locally finite fashion. The very same can also be done in the present case. We preferred not to specify it in the claims, since on the one hand this generalisation makes the claims much less clear at first sight, and on the other hand it is a trivial generalisation once the proof has been completed.

The plan of the paper is as follows. In Section 1.1 we give a short sketch of the proof of Theorem A, to give an idea of how the construction works and to explain where does the constant β in (1.2) come from. In Section 1.2 we recall some basic definitions and properties of sets of finite perimeter. In Section 2 we shall prove Theorem A, which we exploit in Section 3 to prove Theorem B and Theorem C.

1.1 A Quick Sketch of the Proof of Theorem A

Among the three main theorems of this paper, the first one is by far the hardest, and its proof is quite technical. Nevertheless, the overall idea is quite simple, and we outline it in this short section. In particular, the meaning of the value of β in (1.2) will appear as natural. Let us consider a very simplified situation, namely, we assume that the set E is smooth. Of course, the main difficulty of the real proof is exactly to make everything precise also when dealing with non-smooth parts of the boundary.

Figure 1

Sketch of the proof of Theorem A.

Let x be a point of E, and let us assume to fix the ideas that the outer normal vector to E at x is vertical. Let a be a very small quantity, and let us define F by translating vertically of a length δ the part of E which is at distance less than a from x, as depicted in Figure 1. Since the difference between the volumes of E and F must be ε, and since the density f is locally bounded above by hypothesis and below by positivity and l.s.c., we obtain that

(1.5) a N - 1 δ ε .

Let us then evaluate the difference between the perimeters of E and F. The boundary of F coincides with the boundary of E except for a “vertical part” (the two vertical segments in the figure) and for the fact that a “horizontal” piece of the boundary has been translated. The extra part consists of two segments of length |δ| in dimension N=2, while in general it is the lateral boundary of a cylinder of radius a and height |δ|, hence the perimeter behaves like aN-2|δ|. Finally, the part which is translated has (N-1)-dimensional measure of order aN-1. Since it has been vertically translated of a length δ, so in particular the normal vector remains the same, the α-Hölder property of h in the spatial variable ensures that the difference in perimeter is at most of order aN-1|δ|α. Hence, up to a multiplicative constant we can estimate

P ( F ) - P ( E ) a N - 2 | δ | + a N - 1 | δ | α .

Optimising the choice of a and δ subject to the constraint (1.5), we select a=|ε|γ with

γ = 1 - α α + N ( 1 - α ) ,

from which the above estimate becomes precisely P(F)-P(E)|ε|β with β given by (1.2).

1.2 Some Properties of Sets of Locally Finite Perimeter

In this short section we recall some basic properties of sets of finite perimeter. A complete reference for the subject is for instance the book [6], however the few things listed below make the present paper self-contained. In this section, whenever we write perimeter we mean the Euclidean one, while by volume the standard N-dimensional Lebesgue measure.

We say that a Borel set EN is a set of locally finite perimeter if its characteristic function χE is a BVloc function. The reduced boundary*E of a Borel set EN is the collection of points xN such that a direction ν(x)𝕊N-1 exists (and then it is necessarily unique) such that, calling

B r ± ( x ) = { y N : | y - x | < r , ( y - x ) ν ( x ) 0 } ,

one has

lim r 0 | B r + ( x ) E | Eucl | B r + ( x ) | Eucl = 0 , lim r 0 | B r - ( x ) E | Eucl | B r - ( x ) | Eucl = 1 .

We call the direction ν(x) the outer normal of E at x. A remarkable fact is that E has locally finite perimeter if and only if *E has locally finite (N-1)-dimensional Hausdorff measure N-1. Moreover, for any set E of locally finite perimeter one has

μ E := D χ E = - ν ( x ) N - 1 * E .

In particular, one defines the perimeter of any set by PEucl(E)=N-1(*E)=|μE|(N), and a set has locally finite perimeter if and only if for any ball B, EB has finite perimeter. Obviously, whenever E is regular enough, this notion of perimeter coincides with the classical one, and the reduced boundary *E coincides with the topological boundary E up to N-1-negligible subsets.

Another useful characterisation of the boundary is the following. We say that a set EN has density d at a point xN if

lim r 0 | B r + ( x ) E | Eucl | B r + ( x ) | Eucl = d ,

and we call Ed the collection of all the points with density d. By Lebesgue Theorem, E=E1 and NE=E0 up to N-negligible subsets, so that N(N(E0E1))=0. A stronger fact holds true, namely, a set E has finite perimeter if and only if N-1(N(E0E1)) is finite, and moreover the three sets *E,E12, and N(E0E1) coincide up to N-1-negligible subsets. Thus, the (N-1)-dimensional Hausdorff measure of each of these can be equivalently used as a generalised notion of perimeter.

Let us now conclude this section by listing two fundamental, well-known results about sets of locally finite perimeter.

Theorem 2 (Blow-Up).

Let ERN and let x*E. For every ε>0, let Eε:=1ε(E-x) be the blow-up set at x, write for the sake of brevity με:=μEε and call Hx={yRN:yν(x)=0} the tangent hyper-plane to E at x, and Hx-={yRN:yν(x)<0}. Then as ε0+ the sets Eε converge in the Lloc1-sense to the half-space Hx-, while the measures με (resp. |με|) converge in the weak* sense to the measure νE(x)HN-1Hx- (resp. HN-1Hx-).

Definition 3 (Vertical and Horizontal Sections).

Given a Borel set EN, we define its vertical section at level yN-1, Ey, and its horizontal section at level t, Et, respectively as

E y := { t : ( y , t ) E } , E t := { y N - 1 : ( y , t ) E } .

The following theorem, which goes by the name of Vol’pert Theorem, states that almost all (with respect to the proper dimensional Hausdorff measure) vertical sections and horizontal sections of sets of locally finite perimeter are of finite perimeter. Moreover, the reduced boundary of the sections coincides with the sections of the reduced boundary. The proof for vertical sections can be found in [6, 35], while for horizontal ones in [7, 23, 24].

Theorem 4 (Vol’pert).

Let E be a set of locally finite perimeter. Then for HN-1-a.e. yRN-1 the vertical section Ey is a set of finite perimeter in R, and *(Ey)=(*E)y. Analogously for the horizontal sections, i.e. for H1-a.e. tR the horizontal section Et is a set of finite perimeter in RN-1, and *(Et)=(*E)t up to an Hn-2-negligible set.

To fully understand the meaning of the Vol’pert Theorem, it is useful to consider what follows. The set *E is an (N-1)-dimensional set, so for 1-almost every t the section (*E)t is well defined up to N-2-negligible subsets. Concerning *(Et), this is well-defined whenever the section Et is a well-defined (N-1)-dimensional set. However since E has locally finite perimeter, N-1-a.e. point of N has density either 1, or 0, or 12. Moreover, the points of density 12 have locally finite N-1-measure. Consequently, only countably many sections Et carry a strictly positive N-1-measure of points with density 12. Hence, for a.e. t, N-1-almost every point of N, whose last coordinate equals t, has either density 0 or 1. For these t, the section Et is univocally defined, and thus the claim of Vol’pert Theorem makes perfect sense. A fully analogous consideration holds for vertical sections.

2 Proof of the ε-εβ Property

This whole section is devoted to the proof of the ε-εβ property. This is quite involved, thus for the sake of clarity we split it in several steps.

Proof of Theorem A.

Let EN be a set of locally finite perimeter, and let BN be a ball such that N-1(B*E)>0. Since f and h are locally bounded and l.s.c., there exists some constant M>0 such that

1 M f ( y ) M , 1 M h ( y , ν ) M    for all  y B , ν 𝕊 N - 1 .

Let x*EB be any point of the reduced boundary of E in B. Up to a rotation and a translation, we may assume that x=0 and that νE(x)=(0,1)N-1×.

Step I: Choice of the reference cube and the “good” part G. First of all, we let ρ=ρ(M,N)>0 be a small parameter, only depending on M and on N, which will be chosen later on. As a direct application of the Blow-Up Theorem 2 we obtain the existence of an arbitrarily small constant a>0 such that, calling QN=(-a2,a2)N and Q=(-a2,a2)N-1 the N-dimensional and the (N-1)-dimensional cubes of side a, and letting x=(x,xN)N-1× one has

(2.1) 1 - ρ N - 1 ( * E Q N { - a ρ < x N < a ρ } ) a N - 1 1 + ρ ,
(2.2) N - 1 ( * E Q N { - a ρ < x N < a ρ } ) ρ a N - 1 ,
(2.3) N ( ( E { x N < 0 } ) Q N ) < ρ 2 a N ,
(2.4) N - 2 ( * E Q N ) 2 ( N - 1 ) a N - 2 1 + ρ .

In fact, the Blow-Up Theorem ensures the first three properties for every small a, and the validity of the last one for some arbitrarily small value of a is then clear by integration. Notice that a depends only on ρ and on E, so ultimately a=a(M,N,E). Observe that, again in view of the Blow-Up Theorem 2, inside the cube QN one has to expect the set E to be very close to the lower half-cube. As a consequence, a “standard” vertical section of E inside QN should be close to the segment (-a2,0). Writing then for brevity ExQ=ExQN and *ExQ=*ExQ, we define then GQ the “good” sections, that is,

G := { x Q : * ( E x ) = ( * E ) x , # ( * E x Q ) = 1 , * E x Q ( - a ρ , a ρ ) , E x Q ( - a 2 , a ρ ) } .

We want to prove that the set G covers most of the cube Q, and actually only very few perimeter is carried by sections which are not in G. More precisely, we will prove that

(2.5) N - 1 ( Q G ) 4 ρ a N - 1 ,
(2.6) N - 1 ( * E ( ( Q G ) × ( - a 2 , a 2 ) ) ) 6 ρ a N - 1 .

To obtain these estimates, let us consider xQ such that xG. This can happen for different reasons, for instance x can belong to the set Γ1 of the sections for which *(Ex)(*E)x, or to the set Γ2 of the sections such that (*E)x contains at least two points. Let us then call Γ3=Q(GΓ1Γ2), and consider a point x in Γ3. We have that *(Ex)=(*E)x and that it contains either no point, or a single point. In this second case, this single point either does not belong to (-aρ,aρ), or it belongs to (-aρ,aρ) and the section Ex is the upper part of the segment instead of the lower one, that is, Ex(-aρ,a2). In any case, we immediately obtain that for any xΓ3 one has

1 ( E x ( - a 2 , 0 ) ) a ρ ,

and by (2.3) we deduce

(2.7) N - 1 ( Γ 3 ) ρ a N - 1 .

Instead, Vol’pert Theorem 4 gives that

(2.8) N - 1 ( Γ 1 ) = 0 .

Since the projection on the first N-1 coordinates is 1-Lipschitz and by definition of Γ2 we have

(2.9) N - 1 ( * E ( G × ( - a 2 , a 2 ) ) ) N - 1 ( G ) ,
N - 1 ( * E ( Γ 2 × ( - a 2 , a 2 ) ) ) 2 N - 1 ( Γ 2 ) .

Therefore, on the one hand by (2.8) and (2.7) we get

N - 1 ( * E Q N ) N - 1 ( G ) + 2 N - 1 ( Γ 2 ) = a N - 1 - N - 1 ( Γ 1 ) - N - 1 ( Γ 3 ) + N - 1 ( Γ 2 ) ( 1 - ρ ) a N - 1 + N - 1 ( Γ 2 ) .

On the other hand, (2.1) and (2.2) give

(2.10) N - 1 ( * E Q N ) ( 1 + 2 ρ ) a N - 1 ,

so we deduce

N - 1 ( Γ 2 ) 3 ρ a N - 1 .

Since Q=G(Γ1Γ2Γ3), this estimate together with (2.8) and (2.7) implies (2.5). Finally, (2.5) together with (2.9) and (2.10) directly gives (2.6).

Step II: An estimate about small cubes. In this step we prove a simple estimate about the perimeter on small cubes. More precisely, for every xN-1 and >0 let us call Q(x)N-1 the cube with center in x, side , and with sides parallel to the coordinate planes. Let us then fix some cQ and some >0 such that Q2(c)Q. We aim to show that

(2.11) Q ( c ) N - 2 ( * E ( Q ( x ) × ( - a 2 , a 2 ) ) ) 𝑑 N - 1 ( x ) ( N - 1 ) N - 2 N - 1 ( * E ( Q 2 ( c ) × ( - a 2 , a 2 ) ) ) .

To prove this estimate, let us fix a direction 1iN-1, and for every xQ(c) let us call

S i - ( x ) = { y Q 2 ( c ) : y i = x i - } .

We have then

Q ( c ) N - 2 ( * E ( S i - ( x ) × ( - a 2 , a 2 ) ) ) 𝑑 N - 1 ( x ) = N - 2 - 3 2 - 2 N - 2 ( * E ( Q 2 ( c ) × ( - a 2 , a 2 ) ) { x i = c i + t } ) d t N - 2 N - 1 ( * E ( Q 2 ( c ) × ( - a 2 , a 2 ) ) { c i - 3 2 < x i < c i - 2 } ) .

The same estimate is clearly valid replacing Si- with Si+, defined in the same way with yi=xi+ in place of yi=xi-. Since for every xQ(c) one clearly has Q(x)i=1N-1Si-(x)Si+(x), summing the last estimate among all 1iN-1, we immediately get (2.11).

Step III: Selection of “good” horizontal cubes Qj. We now fix 0γ<1N-1 with γ=γ(N,α), whose precise value is given in (2.39). Moreover, we fix a very small constant ε¯>0, which will be precised later, and which depends on a, M and γ, so that in the end ε¯=ε¯(M,N,E,α). Let also 0<ε<ε¯ be given. In this step, we define H=H(γ,N) pairwise disjoint cubes QjQ, and in the next step we will select one of them to go on with the construction. If γ=0, we only take one cube, that is H(0,N)=1 and the unique cube is Q1=Q itself. If γ>0, instead, we take pairwise disjoint (N-1)-dimensional open cubes Q2(xj+)Q, where we write for brevity =aεγ. Notice that it is possible to find 2H of these pairwise disjoint cubes with

(2.12) H 1 2 N + 1 ε γ ( N - 1 ) .

For each cube, we can find a point xjQ(xj+) such that, calling Qj=Q(xj),

N - 2 ( * E ( Q j × ( - a 2 , a 2 ) ) ) Q ( x j + ) N - 2 ( * E ( Q ( x ) × ( - a 2 , a 2 ) ) ) 𝑑 N - 1 ,

which by (2.11) gives

N - 2 ( * E ( Q j × ( - a 2 , a 2 ) ) ) N - 1 a ε γ N - 1 ( * E ( Q 2 ( x j + ) × ( - a 2 , a 2 ) ) ) .

Keeping in mind that the cubes Q2(xj+) are pairwise disjoint and contained in Q, as well as (2.10), we deduce that for all 1jH,

(2.13) N - 2 ( * E ( Q j × ( - a 2 , a 2 ) ) ) N 2 N + 1 ( a ε γ ) N - 2 ,

provided that ρ1N. In addition, by Vol’pert Theorem 4, we can also assume that for every 1jH,

(2.14) * E ( Q j × ( - a 2 , a 2 ) ) = * ( E ( Q j × ( - a 2 , a 2 ) ) ) N - 2 -a.e.

Notice that E(Qj×(-a2,a2)) is contained in the union of 2(N-1) hyperplanes of dimension N-1, so the boundary in the right-hand side above has to be intended as the corresponding union of the (N-2)-dimensional boundaries. Observe that also in the case γ=0 the unique cube Q1=Q satisfies (2.13), which is in fact weaker than (2.4), and without loss of generality we can also assume the validity of (2.14) by Vol’pert Theorem.

Step IV: Choice of one of the horizontal cubes Qj. In this step we select a particular horizontal cube among those defined in the previous step. In particular, we aim to find some 1jH such that the cubes Qε=Qj and QεN=Qj×(-a2,a2) satisfy

(2.15) N - 1 ( * E Q ε N ) ( a ε γ ) N - 1 1 + 2 N + 5 ρ ,
(2.16) N - 1 ( * E Q ε N { - a ρ < x N < a ρ } ) ( a ε γ ) N - 1 2 N + 3 ρ ,
(2.17) N ( E Q ε N { x N > 0 } ) a N ε γ ( N - 1 ) 2 N + 3 ρ 2 ,
(2.18) N - 1 ( Q ε G ) ( a ε γ ) N - 1 2 N + 5 ρ ,

being G the set defined in Step I. Let us start noticing that everything is trivial for the special case γ=0. Indeed, in this case there is nothing to choose since there is only one cube Q1=Q, hence we only have to check the validity of properties (2.15)–(2.18). And in turn, estimates (2.1), (2.2), (2.3) and (2.5) exactly provide the validity of these four inequalities, with even better constants. By the way, in this particular case the cubes Qε=Q and QεN=QN actually do not even depend on ε.

We focus then on the non-trivial case γ>0. First of all, keeping in mind (2.2) and that the cubes Qj are pairwise disjoint we obtain that strictly less than 14 of the H cubes Qj are those for which

N - 1 ( * E Q ε N { - a ρ < x N < a ρ } ) > 4 ρ a N - 1 H ,

hence by (2.12) for more than 34 of the cubes (2.16) is valid. With the very same argument one obtains, for more than 34 of the cubes Qj, also the validity of (2.17) from (2.3) and of (2.18) from (2.5).

The argument to obtain (2.15) is slightly more involved. More precisely, calling A the projection over Q of *EQN, for every Borel set VQ we define

ζ ( V ) = N - 1 ( * E ( V × ( - a 2 , a 2 ) ) - N - 1 ( V A ) .

It is immediate to notice that ζ is a positive measure and, observing that QAΓ3, from (2.7) and (2.10) we get ζ(Q)3ρaN-1. Hence, arguing as for the other inequalities, we obtain that for more than 34 of the H cubes Qj one has

ζ ( Q j ) 12 ρ a N - 1 H 2 N + 5 ρ ( a ε γ ) N - 1 ,

which implies (2.15). Summarising, we can find at least one index j with 1jH such that all estimates (2.15)–(2.18) hold true for such an index. For future use we remark that, putting together (2.15) and (2.18), we have

(2.19) N - 1 ( * E ( ( Q ε G ) × ( - a 2 , a 2 ) ) ) 2 N + 6 ρ ( a ε γ ) N - 1 .

Step V: Definition of σ± and of the modified sets FδRN. In this step we define a one-parameter family of sets Fδ; in the next step we will select a suitable δ such that F=Fδ satisfies the volume constraint in (1.1), and then we will check that this set F also satisfies the perimeter constraint in (1.1). To present the construction, we need yet another constant δ¯a, which depends on a,M,γ and ε, so ultimately δ¯=δ¯(M,N,E,α,ε). The precise value of δ¯ is given in (2.28). We define the integer

K := a 3 δ ¯ ,

and we take K pairwise disjoint open strips Si=Qε×(σi,σi+δ¯)QεN, with constants aρ<σi<a2-δ¯ for 1iK. Notice that this is possible in view of the definition of K, in particular we can do this in such a way that also the closures Si¯ of the strips are pairwise disjoint. Then by (2.16) and (2.17)

i = 1 K a N - 1 ( * E S i ¯ ) + N ( E S i ) a N - 1 ( * E ( Q ε × ( a ρ , a 2 ) ) + N ( E ( Q ε × ( a ρ , a 2 ) ) 2 N + 3 ρ a N ε γ ( N - 1 ) + 2 N + 3 ρ 2 a N ε γ ( N - 1 ) 2 N + 4 ρ a N ε γ ( N - 1 ) .

Then there exists at least one strip S+:=Si¯, such that

(2.20) a N - 1 ( * E S + ¯ ) + N ( E S + ) 2 N + 4 ρ a N ε γ ( N - 1 ) K 2 N + 6 δ ¯ ρ ( a ε γ ) N - 1 .

For the sake of notation, we write σ+=σi¯, so that S+=Qε×(σ+,σ++δ¯) and we keep in mind that aρ<σ+<a2-δ¯.

On the other hand, by (2.16) we can estimate

2 N + 3 ρ N - 1 ( * E Q ε N { x N < - a ρ } ) ( a ε γ ) N - 1 - a 2 - a ρ N - 2 ( ( * E Q ε N ) t ) ( a ε γ ) N - 1 𝑑 t
= ( 1 2 - ρ ) - a 2 - a ρ N - 2 ( * ( E t ) Q ε ) a N - 2 ε γ ( N - 1 ) 𝑑 t ,

where the last equality follows from Vol’pert Theorem. In particular, for almost every -a2<t<-aρ one has that N-1-a.e. point of the hyper-plane N-1×{t} has density either 0 or 1, and

( * E Q ε N ) t = * ( E t Q ε ) .

Consequently, we can find another section -a2<σ-<-aρ for which in addition

(2.21) N - 2 ( * ( E σ - ) Q ε ) 3 2 N + 3 ρ a N - 2 ε γ ( N - 1 ) .

For every 0<δδ¯ we define the set Fδ as

F δ := E ( Q ε × ( σ - , σ + + δ ) ) ( ( E σ - Q ε ) × ( σ - , σ - + δ ) ) { ( x , x N + δ ) : ( x , x N ) E ( Q ε × ( σ - , σ + ) ) } .

It is easy but important to observe how the set Fδ is defined. The difference between E and Fδ is only inside the cylinder Qε×(σ-,σ++δ). In this cylinder, the flat basis Eσ-Qε is stretched vertically of a height δ, the high central part E(Qε×(σ-,σ+)) is translated vertically of δ, and the short upper part E(Qε×(σ+,σ++δ))ES+ is simply removed. Keep in mind that E has density either 0 or 1 at N-1-a.e. point of the basis Qε×{σ-}, therefore the vertical stretch of Eσ-Qε makes sense.

Step VI: Volume evaluation and choice of the competitor F. In this step we estimate the volume of the sets Fδ defined in the previous step. We will then select a particular one of these sets, which we will simply denote by F, such that

(2.22) | F | = | E | + ε .

We shall then show that such a set is the one required by the ε-εβ property, by checking the validity of the inequality P(F)P(E)+Cεβ.

For a given 0<δ<δ¯, let us define

E + = F E , E 1 - = ( E F ) ( Q ε × ( σ - , σ + ) ) , E 2 - = ( E F ) ( Q ε × ( σ + , σ + + δ ) ) ,

so that

(2.23) | F | = | E | + | E + | - | E 1 - | - | E 2 - | .

The last term is the easiest to estimate. Indeed, since of course E2-E(Qε×(σ+,σ++δ))ES+, recalling that 1M<f,h<M in QN, by (2.20) we directly have

(2.24) | E 2 - | | E S + | M N ( E S + ) 2 N + 6 M δ ¯ ρ ( a ε γ ) N - 1 .

Let us now pass to estimate |E+| and |E1-|. To do so, it is convenient to consider separately the vertical sections corresponding to some xG and the other ones. We start by picking some xQεG. Then (EQε)x is a vertical segment of the form (-a2,y) with some -aρ<y<aρ, so in particular σ-<y<σ+. Therefore, the section (E1-)x is empty, while the section (E+)x is the segment (y,y+δ). By Fubini and (2.18), we have then

(2.25)

| E 1 - ( G × ( - a 2 , a 2 ) ) | = 0 ,
δ ( a ε γ ) N - 1 M ( 1 - 2 N + 5 ρ ) | E + ( G × ( - a 2 , a 2 ) ) | M δ ( a ε γ ) N - 1 .

Let us now instead take xQεG, and assume that (x,xN)E+ for some xN(σ-,σ++δ). The fact that (x,xN)E+=FE implies that (x,xN)E. Instead, the fact that (x,xN)F implies that (x,y)E, where y=max{xN-δ,σ-}. As a consequence, the segment {x}×(xN-δ,xN) intersects *(Ex). Analogously, if (x,xN)E1-, this means that (x,xN)E while (x,max{xN-δ,σ-})E, so that {x}×(xN-δ,xN) intersects again *(Ex). Keeping in mind Vol’pert Theorem, we deduce that

( E + E 1 - ) ( ( Q ε G ) × ( - a 2 , a 2 ) ) τ ( * E ( Q ε G × ( - a 2 , a 2 ) ) ) ,

where for every set AN we define

(2.26) τ ( A ) = { ( x , x N + t δ ) : ( x , x N ) A ,  0 t 1 } .

Therefore, also by (2.19) we deduce

(2.27) | ( E + E 1 - ) ( ( Q ε G ) × ( - a 2 , a 2 ) ) | M N ( ( E + E 1 - ) ( ( Q ε G ) × ( - a 2 , a 2 ) ) ) M δ N - 1 ( * E ( Q ε G × ( - a 2 , a 2 ) ) ) 2 N + 6 M δ ρ ( a ε γ ) N - 1 .

We are finally in a position to define δ¯ as

(2.28) δ ¯ = 2 M ε ( a ε γ ) N - 1 .

Keep in mind that our construction makes sense only if δ¯a, which in turn is true as soon as we have chosen ε¯ small enough (recall that ε¯ could be chosen depending on a,M and γ, and that 0γ<1N-1). Putting together (2.23), (2.24), (2.25) and (2.27), and recalling (2.28), in the case δ=δ¯ we can now estimate

(2.29) | F δ ¯ | - | E | δ ¯ ( a ε γ ) N - 1 ( 1 - 2 N + 5 ρ M - 2 N + 7 M ρ ) = 2 M ε ( 1 - 2 N + 5 ρ M - 2 N + 7 M ρ ) > ε ,

where the last inequality is true up to taking ρ small enough depending on M and N (keep in mind that ρ was chosen precisely depending on M and on N). On the other hand, in the case δ=δ¯4M2, (2.23), (2.25) and (2.27), together with (2.28), allow to deduce

(2.30) | F δ ¯ 4 M 2 | - | E | δ ¯ 4 M 2 ( a ε γ ) N - 1 ( M + 2 N + 6 M ρ ) = ε 2 M ( M + 2 N + 6 M ρ ) < ε ,

again up to choosing ρ small enough. Since of course the measure of Fδ is continuous on δ, by (2.29) and (2.30) we deduce the existence of some δ between δ¯4M2 and δ¯ such that (2.22) holds. We fix then this δ, and from now on we let F=Fδ.

Step VII: Evaluation of P(F). In this step we evaluate P(F). Notice that, in order to obtain the ε-εβ property, we need to show that P(F)P(E)+Cεβ. Since F equals E outside of the open cylinder 𝒞=Qε×(σ-,σ++δ), the difference between *F and *E is contained in the closed cylinder. We split *F𝒞¯ in four parts, namely Ttop+, Tbtm+, Tltr+ and Tint+, where

(2.31)

T top + = * F ( Q ε × { σ + + δ } ) , T btm + = * F ( Q ε × { σ - } ) ,
T ltr + = * F ( Q ε × ( σ - , σ + + δ ) ) , T int + = * F 𝒞 .

We also split *E𝒞¯ in four parts in a slightly different way as

(2.32)

T top - = * E ( Q ε × [ σ + , σ + + δ ] ) , T btm - = * E ( Q ε × { σ - } ) ,
T ltr - = * E ( Q ε × ( σ - , σ + + δ ) ) , T int - = * E ( Q ε × ( σ - , σ + ) ) .

We now have to carefully consider the above pieces. The easiest thing to notice is that, since by definition of σ- the set E has density either 0 or 1 at N-1-almost every point of Qε×{σ-}, then

(2.33) N - 1 ( T btm + T btm - ) = 0 ,

so neither E nor F carry any perimeter on the bottom face of the cylinder 𝒞.

Let us now consider Ttop±. We aim to show that, up to N-1-negligible subsets,

(2.34) T top + π ( T top - ) ,

where π(x,xN)=(x,σ++δ) denotes the projection over the hyperplane {xN=σ++δ}. By the properties of sets of finite perimeter, for N-1-almost every xQε we have that E has density either 0, or 12, or 1 both at (x,σ+) and at (x,σ++δ). If at least one of them is 12, then x belongs to the right set in (2.34). If they are both 0 or both 1, then by construction F has also density 0 or 1 at the point (x,σ++δ), which then does not belong to Ttop+. Let us then assume that one of the two densities is 0 and the other one is 1. Up to discarding an N-1-negligible quantity of points x, we can assume that Vol’pert Theorem holds for the section Ex, and the density of Ex at the points σ+ and σ++δ is once 0 and once 1. Consequently, *(Ex) contains some point in the segment (σ+,σ++δ), and then again x belongs to the right set in (2.34). Summarising, we proved the inclusion (2.34) up to sets of zero N-1-measure. Keeping in mind the fact that the projection is 1-Lipschitz and (2.20), and noticing that Ttop-*ES+¯, we have then

(2.35) T top + h ( y , ν F ( y ) ) 𝑑 N - 1 ( y ) M N - 1 ( T top + ) M N - 1 ( T top - ) 2 N + 6 M δ ¯ ρ a N - 2 ( ε γ ) N - 1 .

To consider the set Tltr+, we can argue similarly, keeping in mind that by (2.14) Vol’pert Theorem is true for N-2-almost each xQε. Indeed, take (x,xN) where x is a point in Qε at which Vol’pert Theorem holds true, and σ-+δ<xN<σ++δ. Up to N-1-negligible subsets, E has density either 0, or 12 or 1 both at (x,xN) and at (x,xN-δ). If the densities are both 0 or both 1, then (x,xN)*F, while if one density is 0 and the other one is 1 then again there is some point in the segment {x}×(xN-δ,xN) which belongs to *E. Summarising, up to N-1-negligible subsets

T ltr + Q ε × ( σ - , σ - + δ ) τ ( * E ( Q ε × ( σ - + δ , σ + + δ ) ) ) ,

where τ is defined in (2.26). As a consequence, also recalling (2.13) we have

(2.36) T ltr + h ( y , ν F ( y ) ) 𝑑 N - 1 ( y ) M ( N - 1 ( Q ε × ( σ - , σ - + δ ) ) + δ N - 2 ( * E ( Q ε × ( - a 2 , a 2 ) ) ) ) M ( 2 δ ( N - 1 ) ( a ε γ ) N - 2 + N δ 2 N + 1 ( a ε γ ) N - 2 ) 2 N + 2 N M δ ( a ε γ ) N - 2 .

To conclude, we have to consider Tint±. Since the section σ- has been chosen in such a way that Vol’pert Theorem is true, and in the cylinder 𝒞-=Qε×(σ-,σ-+δ) the set F is simply (Eσ-Qε)×(σ-,σ-+δ), we readily obtain

* F 𝒞 - = ( * ( E σ - ) Q ε ) × ( σ - , σ - + δ ) .

Instead, by construction, *F(Qε×(σ-+δ,σ++δ)) is nothing else than a vertical translation of height δ of the set *E(Qε×(σ-,σ+))=Tint-. In particular, if (x,xN)Tint-, then νF(x,xN+δ)=νE(x,xN). Let us call for brevity ξ the vertical translation of a height δ, i.e. ξ(x,xN)=(x,xN+δ), so that for yTint- one has νF(ξ(y))=νE(y). By the local α-Hölder assumption on h in the spatial variable, we have a constant C1 such that |h(ξ(y),ν)-h(y,ν)|C1δα for every yQN, ν𝕊N-1. Consequently, by (2.21) and (2.15) we can estimate

T int + h ( y , ν F ( y ) ) 𝑑 N - 1 ( y ) - T int - h ( y , ν E ( y ) ) 𝑑 N - 1 ( y )
M N - 1 ( * F 𝒞 - ) + T int - h ( ξ ( y ) , ν E ( y ) ) - h ( y , ν E ( y ) ) d N - 1 ( y )
M δ N - 2 ( * ( E σ - ) Q ε ) + C 1 δ α N - 1 ( T int - )
(2.37) 3 2 N + 3 M δ ρ a N - 2 ε γ ( N - 1 ) + C 1 δ α ( a ε γ ) N - 1 ( 1 + 2 N + 5 ρ ) .

Putting together (2.33), (2.35), (2.36) and (2.37), and recalling that δδ¯ together with the definition (2.28) of δ¯, if ε¯ is small enough, we finally derive

(2.38) P ( F ) - P ( E ) C 2 ( δ ¯ ( ε γ ) N - 2 + δ ¯ α ( ε γ ) N - 1 ) C 3 ( ε 1 - γ + ε α + γ ( N - 1 ) ( 1 - α ) )

for two constants C2,C3 only depending on N,M and a, so actually on N,M and E.

Step VIII: Optimal choice of γ and definition of β. Estimate (2.38) proved in the above step holds for a generic γ. Keeping in mind that γ could be chosen depending only on N and α, and that the construction requires 0γ<1N-1, we here make an optimal choice. Notice that γ1-γ is strictly decreasing while γα+γ(N-1)(1-α) is increasing. As 1-γα+γ(N-1)(1-α) for γ=0 while 1-γ<α+γ(N-1)(1-α) for γ=1N-1, the best choice is the unique value γ such that 1-γ=α+γ(N-1)(1-α), i.e.

(2.39) γ = 1 - α α + N ( 1 - α ) .

With this choice, estimate (2.38) becomes P(F)P(E)+Cεβ where β=β(N,α) is given by (1.2) and C only depends on M,N,E,α and the local Hölder constant of h.

Step IX: The case of a continuous function h. In this step we consider in more details the case α=0, in which “locally α-Hölder” is a synonym of “locally bounded”. In this case, γ=1N and β=N-1N, so in the previous steps for a small 0<ε<ε¯ we already found a set F such that |F|=|E|+ε and P(F)P(E)+CεN-1N. In view of applications, for instance Theorem B, it is important to be able to choose the constant C arbitrarily small. More precisely, for any given constant κ>0, one would be interested to have Cκ up to choosing ε¯ accordingly. In particular, ε¯ should also depend on κ and clearly if κ becomes very small, so must ε¯. It is simple to see that this improvement is false for a generic locally bounded h, see for instance [14]. Here we show that such an improvement is possible if h is continuous in the spatial variable.

Let us assume that α=0 and that h is continuous in the spatial variable. We only need a slight yet fundamental modification of our argument. More precisely, we fix a large constant L=L(M,N,E,κ), to be specified later on, and we possibly decrease, depending on κ and on L, the value of ε¯ found in the previous steps: we let ε¯=ε¯L. For every 0<ε<ε¯, we call ε=εL, so that ε can be any number in the interval (0,ε¯). We aim to find some set F, which equals E outside QN, such that

(2.40) | F | = | E | + ε , P ( F ) P ( E ) + C ε N - 1 N .

Correspondingly, we also modify the definition of δ¯, which in place of (2.28) is now

(2.41) δ ¯ = 2 M ε L ( a ε γ ) N - 1 = 2 M ε 1 N L a N - 1 .

As already done immediately after (2.28), we recall that the construction only makes sense with δ¯a, and this is again true as soon as ε¯ is small enough. Up to these modifications, we keep everything unchanged in the first five steps. In Step VI, the volume estimates are again true, and in particular (2.29) and (2.30) now read

| F δ ¯ | - | E | δ ¯ ( a ε γ ) N - 1 ( 1 - 2 N + 5 ρ M - 2 N + 7 M ρ ) > ε L = ε ,
| F δ ¯ 4 M 2 | - | E | δ ¯ 4 M 2 ( a ε γ ) N - 1 ( M + 2 N + 6 M ρ ) = ε 2 L M ( M + 2 N + 6 M ρ ) < ε L = ε ,

as soon as ρ is small enough and depending only on N and M. Again by continuity, we have then the existence of a constant δ¯4M2<δ<δ¯ such that, calling F=Fδ, the equality |F|-|E|=ε holds. We have then to bound the perimeter of F, and this will be done by using the estimates of Step VII with only a single modification.

Exactly in Step VII, we divide *F𝒞¯ and *E𝒞¯ in the parts Ttop±, Tbtm±, Tltr± and Tint± by (2.31) and (2.32). Recall that N-1(Tbtm+Tbtm-)=0 by (2.33), while by (2.35) and (2.36), also recalling (2.41) and that γ=1N, we have

(2.42) T top + T ltr + h ( y , ν F ( y ) ) 𝑑 N - 1 ( y ) 2 N + 7 N M 2 a L ( ρ ε + ε N - 1 N ) = C 4 ( ρ ε + 1 L 1 N ε N - 1 N ) < κ 3 ε N - 1 N ,

where C4 is a constant only depending on M,N and a, so again on M,N and E, and the last step is true as soon as L is large enough and ε¯ is small enough, both depending on M,N,E and κ.

To conclude, we need to evaluate the perimeter contribution of Tint±. We will argue similarly as what already done in estimate (2.37). The only difference is that this time we do not estimate

h ( ξ ( y ) , ν E ( y ) ) - h ( y , ν E ( y ) ) C 1 δ α = C 1

by using the local boundedness, rather we use the continuity of h in the spatial variable. This implies uniform continuity when the spatial variable is inside the cube QN. More precisely, we call ω the continuity modulus of h in the spatial variable inside QN, that is

ω ( d ) = sup { | h ( y , ν ) - h ( z , ν ) | : y , z Q N , | y - z | d , ν 𝕊 N - 1 } .

Then, calling ξ, as in Step VII, the vertical translation of height δ, and recalling that δ<δ¯ and (2.41), we have

| h ( ξ ( y ) , ν E ( y ) ) - h ( y , ν E ( y ) ) | ω ( δ ¯ ) = ω ( 2 M ε 1 N L a N - 1 ) .

Putting this estimate inside (2.37), we now get

T int + h ( y , ν F ( y ) ) 𝑑 N - 1 ( y ) - T int - h ( y , ν E ( y ) ) 𝑑 N - 1 ( y ) 3 2 N + 3 M δ ¯ ρ a N - 2 ε γ ( N - 1 ) + ω ( δ ¯ ) ( a ε γ ) N - 1 ( 1 + 2 N + 5 ρ )
C 5 ε + 2 ω ( δ ¯ ) a N - 1 L N - 1 N ε N - 1 N
(2.43) < κ 3 ε N - 1 N + 2 ω ( δ ¯ ) a N - 1 L N - 1 N ε N - 1 N ,

where as usual C5=C5(M,N,a) and the last estimate is true for ε¯ small enough, as usual depending on M,N,E and κ.

We can finally conclude. Indeed, the first thing to do is to fix L, depending on M,N,E and κ, so large that (2.42) holds true. Once L has been fixed, we fix ε¯ as small as desired, depending on M,N,E,κ and L. Since h is continuous in the spatial variable, the continuity modulus ω(d) goes to 0 as d goes to 0. Hence, if δ¯ is small enough, we have that

2 ω ( δ ¯ ) a N - 1 L N - 1 N < κ 3 .

Recalling the definition (2.41) of δ¯, we deduce that the above inequality is true as soon as ε¯ is small enough, depending on M,N,L and a. Inserting this last inequality in (2.43) and recalling (2.42), we obtain P(F)P(E)+κεN-1N. Summarising, in the case when α=0 but h is continuous in the spatial variable, for every κ>0 we have found some ε¯>0 depending on M,N,E,κ such that for every 0<ε<ε¯ there is a set F with EFQN satisfying (2.40). The ε-εβ property with constant κ has then been proved, at least for positive values of ε.

Step X: Conclusion (i.e. the case ε<0). In this last step, we finally conclude the proof. We are left to consider the case ε<0, which is in fact a simple consequence of what we have proved so far. Summarising, we have taken a set EN of locally finite perimeter, a ball B with N-1(*EB)>0, and a point x*EB, and we found two positive constants C and ε¯ with the property that, for every 0<ε<ε¯, there exists a set F with FEB and satisfying (1.1), that is

F E B , | F | - | E | = ε , P ( F ) - P ( E ) C ε β .

We can apply this result to the set E^=BE, which is also a set of locally finite perimeter and whose reduced boundary also have intersection with B of strictly positive N-1-measure, since *E^B=*EB. Then we have two positive constants C^ and ε¯^ such that for every 0<ε<ε¯^, there exists a set F^ such that

F ^ E ^ B , | F ^ | - | E ^ | = ε , P ( F ^ ) - P ( E ^ ) C ^ ε β .

Defining then F=(EB)(BF^), we clearly have

F E B , | F | - | E | = - ε , P ( F ) - P ( E ) C | - ε | β .

In other words, we automatically have the validity of (1.1) also for negative values of ε, up to possibly decreasing the value of ε¯ (resp., increasing the value of C), in case ε¯^ is smaller (resp., C^ is larger). The proof is then concluded. ∎

3 Boundedness and Regularity of Isoperimetric Sets

In this last section we give the proofs of Theorem B and of Theorem C, which respectively deal with the boundedness and the regularity of isoperimetric sets.

3.1 Boundedness

Let us start with the proof of the boundedness of isoperimetric sets. We underline that the assumptions of the theorem, namely, the boundedness of the densities and the continuity of h, are both necessary. Indeed, even in the special case of single density, it is possible to find unbounded isoperimetric sets when either the boundedness or the continuity assumption is dropped (see [29, 14]).

Proof of Theorem B.

Let E be an isoperimetric set as in the claim. In particular, we can find a ball B and some ε¯>0 such that for every -ε¯<ε<ε¯ there exists a set F satisfying (1.1) with constant

C = N ω N 1 N 2 M 2 N - 1 N .

Let us fix R01 such that, calling BR0={xN:|x|<R0}, one has BBR0 and |EBR0|<ε¯. For every R>R0, let us define

φ ( R ) = | E B R | ,

which is a bounded and locally Lipschitz decreasing function, thus is particular in W1,1(). For every R>R0, we can take a set F satisfying (1.1) with ε=φ(R), so that |FBR|=|E|. Since E is an isoperimetric set, keeping in mind (1.3) we can evaluate

P ( E ) P ( F B R ) = P ( F ) + P ( E B R ) - P ( E ) P ( F ) - P ( E B R ) + B R E h ( x , x | x | ) + h ( x , - x | x | ) d N - 1 ( x ) P ( E ) + C φ ( R ) N - 1 N - P ( E B R ) + 2 M N - 1 ( B R E ) P ( E ) + C φ ( R ) N - 1 N - P ( E B R ) - 2 M 2 φ ( R ) ,

recalling that

φ ( R ) = - B R E f ( x ) 𝑑 N - 1 ( x )

holds for N-1-almost every R.

By (1.3) and by the Euclidean isoperimetric inequality we have

P ( E B R ) 1 M P Eucl ( E B R ) N ω N 1 N M | E B R | Eucl N - 1 N N ω N 1 N M 2 N - 1 N φ ( R ) N - 1 N = 2 C φ ( R ) N - 1 N .

Pairing it with the previous estimate, we get

| φ ( R ) | C 2 M 2 φ ( R ) N - 1 N .

Since N-1N<1, this inequality implies the existence of some R1>R0 such that φ(R1)=0, that is, the set E is bounded. ∎

3.2 Regularity

To present the proof of the regularity of isoperimetric sets, we first need to recall some classical definitions and results.

Definition 1.

We say that a set EN of locally finite perimeter is quasi-minimal if there exists a constant K>0 such that, for every ball Br(x), one has

(3.1) P Eucl ( E , B r ( x ) ) K r N - 1 .

We say it is ω-minimal, for some continuous and increasing function ω:++ with ω(0)=0, if for every ball Br(x) and every set HN with HEBr(x) one has

P Eucl ( E , B r ( x ) ) P Eucl ( H , B r ( x ) ) + ω ( r ) r N - 1 .

We say that a set EN is porous if there exists a constant δ>0 such that, for every xE and every r>0 small enough (possibly depending on x), there are two balls B1,B2Br(x), both with radius δr, such that B1E and B2NE.

Putting together well-known results, see [15, 26, 27, 34], we have the following.

Theorem 2.

Let E be a set of locally finite perimeter. If it is locally quasi-minimal, then it is porous and E=*E up to HN-1-negligible sets. Additionally, if it is locally ω-minimal with ω(r)=Cr2η for some η(0,12], then E is C1,η.

We are now ready to prove Theorem C.

Proof of Theorem C.

Let EN be an isoperimetric set. Thanks to Theorem 2, we only need to check that E is locally quasi-minimal whenever f and h are locally bounded, and that if in addition h=h(x) is α-Hölder for some 0<α1, then E is also locally ω-minimal with ω(r)=Cr2η, being η as in (1.4).

We fix a generic ball BN, and we only need to consider balls Br(x)B. Since f and h are locally bounded and l.s.c., there exists some constant M>0 such that

1 M f ( x ) M , 1 M h ( x , ν ) M for all  x B , ν 𝕊 N - 1 .

Let us start to consider the quasi-minimality. First of all, by Theorem A we know that the ε-εN-1N property holds for E. Then take two disjoint balls B1 and B2, both intersecting *E in a set of positive N-1-measure, and let C1,C2 and ε¯1,ε¯2 be the corresponding constants according to (1.1). Let also C=max{C1,C2} and ε¯=min{ε¯1,ε¯2}, and let

r ¯ = min { ( ε ¯ M ω N ) 1 N , dist ( B 1 , B 2 ) } .

Let now Br(x)B be any ball. If r<r¯, then by definition of r¯ we have that

(3.2) ε := | B r ( x ) E | < M ω N r N < ε ¯ ,

and that Br(x) cannot intersect both B1 and B2. Thus, without loss of generality we may assume that Br(x)B1=. As a consequence, by the ε-εN-1N property we can find a set FN such that FEB1, |F|=|E|+ε and P(F)P(E)+CεN-1N. Calling then G=FBr(x), we have that |G|=|E|, and by the minimality of E, (1.3) and (3.2) we have

P ( E ) P ( G ) P ( F ) - P ( E , B r ( x ) ) + N ω N M r N - 1 P ( E ) + C ε N - 1 N - P ( E , B r ( x ) ) + N ω N M r N - 1 P ( E ) - P ( E , B r ( x ) ) + N ω N M r N - 1 + C M N - 1 N ω N N - 1 N r N - 1 .

Hence, (3.1) holds true for Br(x) with the choice

K = N ω N M 2 + C M 1 + N - 1 N ω N N - 1 N .

If otherwise rr¯, we clearly have

P Eucl ( E , B r ( x ) ) M P ( E , B r ( x ) ) M P ( E ) M P ( E ) r ¯ N - 1 r N - 1 ,

so that (3.1) again holds true for Br(x). The local quasi-minimality of E is then proved.

Let us then assume that h only depends on the spatial variable and that it is α-Hölder, so that by Theorem A we have the validity of the ε-εβ property with β given by (1.2). To conclude the proof, we have to check the ω-minimality of E for balls Br(x)B with ω(r)=C¯r2η, being η as in (1.4) and being C¯ a suitable constant. Up to increasing the constant C¯, we can restrict ourselves to consider only radii r<r¯, being r¯ as before. Let then Br(x)B be any ball, and HN a set such that HEBr(x). Moreover, let us call ε=|E|-|H|, which satisfies |ε|<ε¯ since r<r¯. Arguing exactly as in the first part of the proof, we can find a set FN such that FE is a positive distance apart from Br(x), that |F|=|E|+ε, and that P(F)P(E)+C|ε|β. Calling G=(FBr(x))(HBr(x)), by construction we have that |G|=|E|, so by the isoperimetric property of E we can evaluate

P ( E ) P ( G ) = P ( F ) + P ( H , B r ( x ) ) - P ( E , B r ( x ) ) P ( E ) + C | ε | β + P ( H , B r ( x ) ) - P ( E , B r ( x ) ) .

Calling then m=max{h(x):xBr(x)}1M, by the α-Hölder property of h and by (3.2) we have

( m - C r α ) P Eucl ( E , B r ( x ) ) P ( E , B r ( x ) ) P ( H , B r ( x ) ) + C | ε | β m P Eucl ( H , B r ( x ) ) + C | ε | β m P Eucl ( H , B r ( x ) ) + C ( M ω N r N ) β .

By the first part of the proof E is quasi-minimal, thus

P Eucl ( E , B r ( x ) ) P Eucl ( H , B r ( x ) ) + C M 1 + β ω N β r N β + C M K r α + N - 1 C ¯ r N β ,

where the last inequality comes from the fact that Nβα+N-1, as one readily obtains from equation (1.2). We established the ω-minimality of E with ω(r)=C¯rNβ-(N-1), which completes the proof since we have Nβ-(N-1)=2η. ∎


Communicated by Ireneo Peral


Funding statement: Both authors are members of the INdAM institute and have been partly supported by the INdAM–GNAMPA 2019 project “Problemi isoperimetrici in spazi Euclidei e non” (project number U-UFMBAZ-2019-000473 11-03-2019).

References

[1] E. Abreu and L. G. Fernandes, Jr., On existence and nonexistence of isoperimetric inequality with differents monomial weights, preprint (2019), https://arxiv.org/abs/1904.01441. 10.1007/s00041-021-09900-8Search in Google Scholar

[2] A. Alvino, F. Brock, F. Chiacchio, A. Mercaldo and M. R. Posteraro, Some isoperimetric inequalities on N with respect to weights |x|α, J. Math. Anal. Appl. 451 (2017), no. 1, 280–318. 10.1016/j.jmaa.2017.01.085Search in Google Scholar

[3] A. Alvino, F. Brock, F. Chiacchio, A. Mercaldo and M. R. Posteraro, On weighted isoperimetric inequalities with non-radial densities, Appl. Anal. 98 (2019), no. 10, 1935–1945. 10.1080/00036811.2018.1506106Search in Google Scholar

[4] A. Alvino, F. Brock, F. Chiacchio, A. Mercaldo and M. R. Posteraro, Some isoperimetric inequalities with respect to monomial weights, preprint (2019), https://arxiv.org/abs/1907.03659. 10.1051/cocv/2020054Search in Google Scholar

[5] A. Alvino, F. Brock, F. Chiacchio, A. Mercaldo and M. R. Posteraro, The isoperimetric problem for a class of non-radial weights and applications, J. Differential Equations 267 (2019), no. 12, 6831–6871. 10.1016/j.jde.2019.07.013Search in Google Scholar

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

[7] M. Barchiesi, F. Cagnetti and N. Fusco, Stability of the Steiner symmetrization of convex sets, J. Eur. Math. Soc. (JEMS) 15 (2013), no. 4, 1245–1278. 10.4171/JEMS/391Search in Google Scholar

[8] W. Boyer, B. Brown, G. R. Chambers, A. Loving and S. Tammen, Isoperimetric regions in n with density rp, Anal. Geom. Metr. Spaces 4 (2016), no. 1, 236–265. 10.1515/agms-2016-0009Search in Google Scholar

[9] F. Brock, F. Chiacchio and A. Mercaldo, An isoperimetric inequality for Gauss-like product measures, J. Math. Pures Appl. (9) 106 (2016), no. 2,375–391. 10.1016/j.matpur.2016.02.014Search in Google Scholar

[10] X. Cabré, X. Ros-Oton and J. Serra, Euclidean balls solve some isoperimetric problems with nonradial weights, C. R. Math. Acad. Sci. Paris 350 (2012), no. 21–22, 945–947. 10.1016/j.crma.2012.10.031Search in Google Scholar

[11] A. Cañete, M. Miranda, Jr. and D. Vittone, Some isoperimetric problems in planes with density, J. Geom. Anal. 20 (2010), no. 2, 243–290. 10.1007/s12220-009-9109-4Search in Google Scholar

[12] C. Carroll, A. Jacob, C. Quinn and R. Walters, The isoperimetric problem on planes with density, Bull. Aust. Math. Soc. 78 (2008), no. 2, 177–197. 10.1017/S000497270800052XSearch in Google Scholar

[13] E. Cinti and A. Pratelli, Regularity of isoperimetric sets in 2 with density, Math. Ann. 368 (2017), no. 1–2, 419–432. 10.1007/s00208-016-1409-ySearch in Google Scholar

[14] E. Cinti and A. Pratelli, The ε-εβ property, the boundedness of isoperimetric sets in N with density, and some applications, J. Reine Angew. Math. 728 (2017), 65–103. 10.1515/crelle-2014-0120Search in Google Scholar

[15] G. David and S. Semmes, Quasiminimal surfaces of codimension 1 and John domains, Pacific J. Math. 183 (1998), no. 2, 213–277. 10.2140/pjm.1998.183.213Search in Google Scholar

[16] G. De Philippis, G. Franzina and A. Pratelli, Existence of isoperimetric sets with densities “converging from below” on N, J. Geom. Anal. 27 (2017), no. 2, 1086–1105. 10.1007/s12220-016-9711-1Search in Google Scholar

[17] A. Díaz, N. Harman, S. Howe and D. Thompson, Isoperimetric problems in sectors with density, Adv. Geom. 12 (2012), no. 4, 589–619. 10.1515/advgeom-2012-0009Search in Google Scholar

[18] L. Di Giosia, J. Habib, L. Kenigsberg, D. Pittman and W. Zhu, Balls isoperimetric in n with volume and perimeter densities rm and rk, preprint (2016), https://arxiv.org/abs/1610.05830. Search in Google Scholar

[19] V. Franceschi and R. Monti, Isoperimetric problem in H-type groups and Grushin spaces, Rev. Mat. Iberoam. 32 (2016), no. 4, 1227–1258. 10.4171/RMI/914Search in Google Scholar

[20] V. Franceschi, A. Pratelli and G. Stefani, On the existence of planar minimizing clusters, to appear. Search in Google Scholar

[21] V. Franceschi and G. Stefani, Symmetric double bubbles in the Grushin plane, ESAIM Control Optim. Calc. Var. 25 (2019), Article ID 77. 10.1051/cocv/2018055Search in Google Scholar

[22] G. Franzina and A. Pratelli, Non-existence of isoperimetric sets in the Euclidean space with vanishing densities, to appear. Search in Google Scholar

[23] N. Fusco, The classical isoperimetric theorem, Rend. Accad. Sci. Fis. Mat. Napoli (4) 71 (2004), 63–107. Search in Google Scholar

[24] N. Fusco, F. Maggi and A. Pratelli, The sharp quantitative isoperimetric inequality, Ann. of Math. (2) 168 (2008), no. 3, 941–980. 10.4007/annals.2008.168.941Search in Google Scholar

[25] N. Fusco, F. Maggi and A. Pratelli, On the isoperimetric problem with respect to a mixed Euclidean–Gaussian density, J. Funct. Anal. 260 (2011), no. 12, 3678–3717. 10.1016/j.jfa.2011.01.007Search in Google Scholar

[26] M. Giaquinta and E. Giusti, Quasiminima, Ann. Inst. H. Poincaré Anal. Non Linéaire 1 (1984), no. 2, 79–107. 10.1016/s0294-1449(16)30429-2Search in Google Scholar

[27] J. Kinnunen, R. Korte, A. Lorent and N. Shanmugalingam, Regularity of sets with quasiminimal boundary surfaces in metric spaces, J. Geom. Anal. 23 (2013), no. 4, 1607–1640. 10.1007/s12220-012-9299-zSearch in Google Scholar

[28] I. McGillivray, An isoperimetric inequality in the plane with a log-convex density, Ric. Mat. 67 (2018), no. 2, 817–874. 10.1007/s11587-018-0382-zSearch in Google Scholar

[29] F. Morgan and A. Pratelli, Existence of isoperimetric regions in n with density, Ann. Global Anal. Geom. 43 (2013), no. 4, 331–365. 10.1007/s10455-012-9348-7Search in Google Scholar

[30] A. Pratelli and G. Saracco, On the isoperimetric problem with double density, Nonlinear Anal. 177 (2018), 733–752. 10.1016/j.na.2018.04.009Search in Google Scholar

[31] A. Pratelli and G. Saracco, Regularity of isoperimetric sets for the two-dimensional isoperimetric problem with double density, to appear. Search in Google Scholar

[32] C. Rosales, A. Cañete, V. Bayle and F. Morgan, On the isoperimetric problem in Euclidean space with density, Calc. Var. Partial Differential Equations 31 (2008), no. 1, 27–46. 10.1007/s00526-007-0104-ySearch in Google Scholar

[33] G. Saracco, Weighted Cheeger sets are domains of isoperimetry, Manuscripta Math. 156 (2018), no. 3–4, 371–381. 10.1007/s00229-017-0974-zSearch in Google Scholar

[34] I. Tamanini, Regularity Results for Almost-minimal Oriented Hypersurfaces in n, Dipartimento di Matematica dell’Università del Salento, Lecce, 1984. Search in Google Scholar

[35] V. A. I. Vol’pert, The spaces BV and quasilinear equations, Math. USSR Sb. 2 (1967), no. 2, 225–267. 10.1070/SM1967v002n02ABEH002340Search in Google Scholar

Received: 2019-11-25
Accepted: 2020-01-03
Published Online: 2020-02-07
Published in Print: 2020-08-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/ans-2020-2074/html
Scroll to top button