This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Plateau borders in soap films
and Gauss’ capillarity theory

Francesco Maggi Department of Mathematics, The University of Texas at Austin, Austin, TX, United States of America maggi@math.utexas.edu Michael Novack Department of Mathematical Sciences, Carnegie Mellon University, Pittsburgh, PA, United States of America mnovack@andrew.cmu.edu  and  Daniel Restrepo Department of Mathematics, Johns Hopkins University, Baltimore, MD, United States of America drestre1@jh.edu
Abstract.

We provide, in the setting of Gauss’ capillarity theory, a rigorous derivation of the equilibrium law for the three dimensional structures known as Plateau borders which arise in “wet” soap films and foams. A key step in our analysis is a complete measure-theoretic overhaul of the homotopic spanning condition introduced by Harrison and Pugh in the study of Plateau’s laws for two-dimensional area minimizing surfaces (“dry” soap films). This new point of view allows us to obtain effective compactness theorems and energy representation formulae for the homotopic spanning relaxation of Gauss’ capillarity theory which, in turn, lead to prove sharp regularity properties of energy minimizers. The equilibrium law for Plateau borders in wet foams is also addressed as a (simpler) variant of the theory for wet soap films.

1. Introduction

1.1. Overview

Equilibrium configurations of soap films and foams are governed, at leading order, by the balance between surface tension forces and atmospheric pressure. This balance is expressed by the Laplace–Young law of pressures, according to which such systems can be decomposed into smooth interfaces with constant mean curvature equal to the pressure difference across them, and by the Plateau laws, which precisely postulate which arrangements of smooth interfaces joined together along lines of “singular” points are stable, and thus observable.

The physics literature identifies two (closely related) classes of soap films and foams, respectively labeled as “dry” and “wet”. This difference is either marked in terms of the amount of liquid contained in the soap film/foam [WH99, Section 1.3], or in terms of the scale at which the soap film/foam is described [CCAE+13, Chapter 2, Section 3 and 4].

In the dry case, Plateau laws postulates that (i) interfaces can only meet in three at a time forming 120-degree angles along lines of “YY-points”; and (ii) lines of YY-points can only meet in fours at isolated “TT-points”, where six interfaces asymptotically form a perfectly symmetric tetrahedral angle; see, e.g. [WH99, Equilibrium rules A1, A2, page 24].

In the wet case, small but positive amounts of liquid are bounded by negatively curved interfaces, known as Plateau borders, and arranged near ideal lines of YY-points or isolated TT-points; see Figure

Refer to caption
(a)(a)(b)(b)
Figure 1.1. (a) A Plateau border develops around a “wet” line of YY-points. The wet region is bounded by interfaces of negative constant mean curvature. The equilibrium condition which needs to hold across the transition lines (here depicted in bold) between the negatively curved interfaces of a Plateau border and the incoming dry interfaces is that these interfaces meet tangentially. In the case of soap films, where the dry interfaces have zero mean curvature, the jump in the mean curvature across the transition lines implies a discontinuity in the gradient of the unit normal. (b) An arrangement of Plateau borders near a tetrahedral singularity. The transition lines are again depicted in bold. The incoming dry interfaces are omitted for clarity.

1.1 and [WH99, Fig. 1.8 and Fig. 1.9]. A “third Plateau law” is then postulated to hold across the transition lines between wet and dry parts of soap films/foams, and can be formulated as follows:

the unit normal to a soap film/foam changes continuously (1.1)
across the transition lines between wet and dry interfaces;\displaystyle\mbox{{\it across the transition lines between wet and dry interfaces}}\,;

see, e.g., [WH99, Equilibrium rule B, page 25] and [CCAE+13, Section 4.1.4]. It is important to recall that Plateau borders play a crucial role in determining the mechanical properties of the many physical and biological systems in which they are observed. As a sample of older and newer papers discussing Plateau borders, we mention here [LL65, JP92, BR97, KHS99, LC99, KHS00, GKJ05, SM15]. Postulate (1.1) is assumed in all these works.

The goal of this paper is answering the natural problem of rigorously deriving the equilibrium condition for Plateau borders (1.1) in the context of Gauss’ capillarity theory. Since the case of soap films is much harder and interesting from the mathematical viewpoint, we will postpone the discussion of foams until the very last section of this introduction. The main highlight is that, in addressing Plateau borders of soap films, we will develop a new “theory of spanning” for surfaces of geometric measure theory (GMT) which will find further applications in the two companion papers [MNR23a, MNR23b]; see the closing of this overview for more details about these additional applications.

We now give an informal description of our approach. The starting point is [MSS19], where the idea is introduced of modeling soap films as regions EE of positive volume |E|=v|E|=v contained in the complement Ω=n+1𝐖\Omega=\mathbb{R}^{n+1}\setminus\mathbf{W} of a “wire frame” 𝐖\mathbf{W} (n=2n=2 is the physical case, although the planar case n=1n=1 is also quite interesting in applications). We associate to EE the surface tension energy n(ΩE)\mathcal{H}^{n}(\Omega\cap\partial E) (where n\mathcal{H}^{n} stands for nn-dimensional (Hausdorff) measure, i.e., area when n=2n=2 and length when n=1n=1), and minimize n(ΩE)\mathcal{H}^{n}(\Omega\cap\partial E) under the constraints that |E|=v|E|=v (for some given v>0v>0) and

ΩE is spanning 𝐖.\mbox{$\Omega\cap\partial E$ is spanning $\mathbf{W}$}\,. (1.2)

From the mathematical viewpoint the meaning assigned to (1.2) is, of course, the crux of the matter. In the informal spirit of this overview, we momentarily leave the concept of “spanning” only intuitively defined.

As proved in [KMS22a], this minimization process leads to the identification of generalized minimizers in the form of pairs (K,E)(K,E) with EΩE\subset\Omega, |E|=v|E|=v, and such that

ΩEK and K is spanning 𝐖.\mbox{$\Omega\cap\partial E\subset K$ and $K$ is spanning $\mathbf{W}$}\,. (1.3)

These pairs are minimizing in the sense that

n(ΩE)+2n(KE)n(ΩE),\mathcal{H}^{n}(\Omega\cap\partial E)+2\,\mathcal{H}^{n}(K\setminus\partial E)\leq\mathcal{H}^{n}(\Omega\cap\partial E^{\prime})\,, (1.4)

whenever EΩE^{\prime}\subset\Omega, |E|=v|E^{\prime}|=v and ΩE\Omega\cap\partial E^{\prime} is spanning 𝐖\mathbf{W}.

If K=ΩEK=\Omega\cap\partial E, then generalized minimizers are of course minimizers in the proper sense. If not, the collapsed interface KEK\setminus\partial E is a surface whose positive area has to be counted with a multiplicity factor 22 (which arises from the asymptotic collapsing along KEK\setminus\partial E of oppositely oriented boundaries in minimizing sequences {Ej}j\{E_{j}\}_{j}, see

Refer to caption
KEK\setminus\partial EEjE_{j}E\partial^{*}ESS𝐖\mathbf{W}
Figure 1.2. Emergence of collapsing along a minimizing sequence {Ej}j\{E_{j}\}_{j} for the minimization of n(ΩE)\mathcal{H}^{n}(\Omega\cap\partial E) among sets EΩ=n+1𝐖E\subset\Omega=\mathbb{R}^{n+1}\setminus\mathbf{W} with |E|=v|E|=v and ΩE\Omega\cap\partial E spanning 𝐖\mathbf{W}, when n=1n=1 and 𝐖\mathbf{W} is the union of three disks in the plane. Notice that for this choice of 𝐖\mathbf{W} the minimization of n(S)\mathcal{H}^{n}(S) among SΩS\subset\Omega such that SS is spanning 𝐖\mathbf{W} is solved by three segments meeting at YY-point. Collapsing is intuitively related to the presence of YY-type and TT-type singularities.

Figure 1.2). We expect collapsing to occur whenever the Plateau problem for 𝐖\mathbf{W} admits one minimizer SS with Plateau-type singularities. Whenever this happens, a wetting conjecture is made: sequences {(Kvj,Evj)}j\{(K_{v_{j}},E_{v_{j}})\}_{j} of generalized minimizers with |Evj|=vj0+|E_{v_{j}}|=v_{j}\to 0^{+} as jj\to\infty will be such that the set of Plateau’s singularities Σ(S)\Sigma(S) of SS is such that sup{dist(x,Evj):xΣ(S)}0\sup\{{\rm dist}(x,E_{v_{j}}):x\in\Sigma(S)\}\to 0. Thus we expect that Plateau’s singularities are never “left dry” in the small volume capillarity approximation of the Plateau problem.

A lot of information about generalized minimizers can be extracted from (1.4), and this is the content of [KMS22a, KMS21, KMS22b]. With reference to the cases when n=1n=1 or n=2n=2, one can deduce from (1.4) that if n(KE)>0\mathcal{H}^{n}(K\setminus\partial E)>0, then KEK\setminus\partial E is a smooth minimal surface (a union of segments if n=1n=1) and that E\partial E contains a regular part E\partial^{*}E that is a smooth constant mean curvature surface (a union of circular arcs if n=1n=1) with negative curvature. This is of course strongly reminiscent of the behavior of Plateau borders, and invites to analyze the validity of (1.1) in this context. A main obstacle is that, due to serious technical issues (described in more detail later on) related to how minimality is expressed in (1.4), it turns out to be very difficult to say much about the “transition line”

EE\partial E\setminus\partial^{*}E

between the zero and the negative constant mean curvature interfaces in KK, across which one should check the validity of (1.1). More precisely, all that descends from (1.4) and a direct application of Allard’s regularity theorem [All72] is that EE\partial E\setminus\partial^{*}E has empty interior in KK. Far from being a line in dimension n=2n=2, or a discrete set of points when n=1n=1, the transition line EE\partial E\setminus\partial^{*}E could very well have positive n\mathcal{H}^{n}-measure and be everywhere dense in KK! With such poor understanding of EE\partial E\setminus\partial^{*}E, proving the validity of (1.1) – that is, the continuity of the unit normals to KEK\setminus\partial E and E\partial^{*}E in passing across EE\partial E\setminus\partial^{*}E – is of course out of question.

We overcome these difficulties by performing a major measure-theoretic overhaul of the Harrison–Pugh homotopic spanning condition [HP16, HP17] used in [MSS19, KMS22a, KMS21, KMS22b] to give a rigorous meaning to (1.2), and thus to formulate the homotopic spanning relaxation of Gauss’ capillarity discussed above.

The transformation of this purely topological concept into a measure-theoretic one is particularly powerful. Its most important consequence for the problem discussed in this paper is that it allows us to upgrade the partial minimality property (1.4) of (K,E)(K,E) into the full minimality property

n(ΩE)+2n(KE)n(ΩE)+2n(KE)\mathcal{H}^{n}(\Omega\cap\partial E)+2\,\mathcal{H}^{n}(K\setminus\partial E)\leq\mathcal{H}^{n}(\Omega\cap\partial E^{\prime})+2\,\mathcal{H}^{n}(K^{\prime}\setminus\partial E^{\prime}) (1.5)

whenever EΩE^{\prime}\subset\Omega, |E|=v|E^{\prime}|=v, ΩEK\Omega\cap\partial E^{\prime}\subset K^{\prime} and KK^{\prime} is spanning 𝐖\mathbf{W}. The crucial difference between (1.4) and (1.5) is that the latter is much more efficient than the former when it comes to study the regularity of generalized minimizers (K,E)(K,E), something that is evidently done by energy comparison with competitors (K,E)(K^{\prime},E^{\prime}). Such comparisons are immediate when working with (1.5), but they are actually quite delicate to set up when we only have (1.4). In the latter case, given a competitor (K,E)(K^{\prime},E^{\prime}), to set up the energy comparison with (K,E)(K,E) we first need to find a sequence of non-collapsed competitors {Ej}j\{E^{\prime}_{j}\}_{j} (with EjΩE^{\prime}_{j}\subset\Omega, |Ej|=v|E^{\prime}_{j}|=v, and ΩEj\Omega\cap\partial E^{\prime}_{j} spanning 𝐖\mathbf{W}) such that n(ΩEj)n(ΩE)+2n(KE)\mathcal{H}^{n}(\Omega\cap\partial E^{\prime}_{j})\to\mathcal{H}^{n}(\Omega\cap\partial E^{\prime})+2\,\mathcal{H}^{n}(K^{\prime}\setminus\partial E^{\prime}). Intuitively, EjE_{j}^{\prime} needs to be a δj\delta_{j}-neighborhood of KEK^{\prime}\cup E^{\prime} for some δj0+\delta_{j}\to 0^{+} and the energy approximation property has to be deduced from the theory of Minkowski content. But applying the theory of Minkowski content to (K,E)(K^{\prime},E^{\prime}) (which is the approach followed, e.g., in [KMS22b]) requires (K,E)(K^{\prime},E^{\prime}) to satisfy rectifiability and uniform density properties that substantially restrict the class of available competitors (K,E)(K^{\prime},E^{\prime}).

In contrast, once the validity of (1.5) is established, a suitable generalization (Theorem 1.2) of the partition theorem of sets of finite perimeter into indecomposable components [ACMM01, Theorem 1] combined with a subtle variational argument (see Figure 1.7) allows us to show that, in any ball BΩB\subset\!\subset\Omega with sufficiently small radius and for some sufficiently large constant Λ\Lambda (both depending just on (K,E)(K,E)), the connected components {Ui}i\{U_{i}\}_{i} of B(KE)B\setminus(K\cup E) satisfy a perturbed area minimizing property of the form

n(BUi)n(BV)+Λ|UiΔV|,\mathcal{H}^{n}(B\cap\partial U_{i})\leq\mathcal{H}^{n}(B\cap\partial V)+\Lambda\,|U_{i}\Delta V|\,, (1.6)

with respect to completely arbitrary perturbations VBV\subset B, VΔUiBV\Delta U_{i}\subset\!\subset B. By a classical theorem of De Giorgi [DG60, Tam84], (1.6) implies (away from a closed singular set of codimension at least 88, which is thus empty if n6n\leq 6) the C1,αC^{1,\alpha}-regularity of BUiB\cap\partial U_{i} for each ii, and thus establishes the continuity of the normal stated in (1.1). In fact, locally at each xx on the transition line, KK is the union of the graphs of two C1,αC^{1,\alpha}-functions u1u2u_{1}\leq u_{2} defined on an nn-dimensional disk, having zero mean curvature above the interior of {u1=u2}\{u_{1}=u_{2}\}, and opposite constant mean curvature above {u1<u2}\{u_{1}<u_{2}\}. We can thus exploit the regularity theory for double-membrane free boundary problems devised in [Sil05, FGS15] to deduce that the transition line EE\partial E\setminus\partial^{*}E is indeed (n1)(n-1)-dimensional, and to improve the C1,αC^{1,\alpha}-regularity of BUiB\cap\partial U_{i} to C1,1C^{1,1}-regularity. Given the mean curvature jump across EE\partial E\setminus\partial^{*}E we have thus established the sharp degree of regularity for minimizers of the homotopic spanning relaxation of Gauss’ capillarity theory.

The measure-theoretic framework for homotopic spanning conditions laid down in this paper provides the starting point for additional investigations that would otherwise seem unaccessible. In two forthcoming companion papers we indeed establish (i) the convergence towards Plateau-type singularities of energy-minimizing diffused interface solutions of the Allen–Cahn equation [MNR23a], and (ii) some sharp convergence theorems for generalized minimizers in the homotopic spanning relaxation of Gauss’ capillarity theory in the vanishing volume limit, including a proof of the above mentioned wetting conjecture [MNR23b].

The rest of this introduction is devoted to a rigorous formulation of the results presented in this overview. We begin in Section 1.2 with a review of the Harrison and Pugh homotopic spanning condition in relation to the classical Plateau problem and to the foundational work of Almgren and Taylor [Alm76, Tay76]. In Section 1.3 we introduce the new measure-theoretic formulation of homotopic spanning and discuss its relation to the measure-theoretic notion of essential connectedness introduced by Cagnetti, Colombo, De Philippis and the first-named author in the study of symmetrization inequalities [CCDPM17, CCDPM14]. In Section 1.4 we introduce the bulk and boundary spanning relaxations of Gauss’ capillarity theory, state a general closure theorem for “generalized soap films” that applies to both relaxed problems (Theorem 1.4). In Section 1.5 we prove the existence of generalized soap film minimizers (Theorem 1.5) and their convergence in energy to solutions to the Plateau problem. A sharp regularity theorem (Theorem 1.6) for these minimizers, which validates (1.1), is stated in Section 1.6. Finally, in Section 1.7 we reformulate the above results in the case of foams, see in particular Theorem 1.7.

1.2. Homotopic spanning: from Plateau’s problem to Gauss’ capillarity

The theories of currents and of sets of finite perimeter, i.e. the basic distributional theories of surface area at the center of GMT, fall short in the task of modeling Plateau’s laws. Indeed, two-dimensional area minimizing currents in 3\mathbb{R}^{3} are carried by smooth minimal surfaces, and thus cannot model YY-type111Currents modulo 33 are compatible with YY-type singularities, but not with TT-type singularities. and TT-type singularities. This basic issue motivated the introduction of Almgren minimal sets as models for soap films in [Alm76]: these are sets Sn+1S\subset\mathbb{R}^{n+1} that are relatively closed in a given open set Ωn+1\Omega\subset\mathbb{R}^{n+1}, and satisfy n(S)n(f(S))\mathcal{H}^{n}(S)\leq\mathcal{H}^{n}(f(S)) whenever f:ΩΩf:\Omega\to\Omega is a Lipschitz (not necessarily injective) map with {fid}Ω\{f\neq{\rm id}\,\}\subset\!\subset\Omega. Taylor’s historical result [Tay76] validates the Plateau laws in this context, by showing that, when222Similar regularity assertions hold when n=1n=1 (by elementary methods) and, in much more recent developments, when n3n\geq 3 [CES22]. n=2n=2, Almgren minimal sets are locally C1,αC^{1,\alpha}-diffeomorphic either to planes, to YY-cones, or to TT-cones.

The issue of proposing and solving a formulation of Plateau’s problem whose minimizers are Almgren minimal sets, and indeed admit Plateau-type singularities, is quite elusive, as carefully explained in [Dav14]. In this direction, a major breakthrough has been obtained by Harrison and Pugh in [HP16] with the introduction of a new spanning condition, which, following the presentation in [DLGM17a], can be defined as follows:

Definition A (Homotopic spanning (on closed sets)).

Given a closed set 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} (the “wire frame”), a spanning class for 𝐖\mathbf{W} is a family 𝒞\mathcal{C} of smooth embeddings of 𝕊1\mathbb{S}^{1} into

Ω=n+1𝐖\Omega=\mathbb{R}^{n+1}\setminus\mathbf{W}

that is closed under homotopies in Ω\Omega, that is, if Φ:[0,1]×SS1Ω\Phi:[0,1]\times\SS^{1}\to\Omega is smooth family of embeddings Φt=Φ(t,):SS1Ω\Phi_{t}=\Phi(t,\cdot):\SS^{1}\to\Omega with Φ0𝒞\Phi_{0}\in\mathcal{C}, then Φt𝒞\Phi_{t}\in\mathcal{C} for every t(0,1]t\in(0,1]. A set SS, contained and relatively closed in Ω\Omega, is said to be 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} if

Sγ,γ𝒞.S\cap\gamma\neq\varnothing\,,\qquad\forall\gamma\in\mathcal{C}\,.

Denoting by 𝒮(𝒞)\mathcal{S}(\mathcal{C}) the class of sets SS 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, one can correspondingly formulate the Plateau problem (with homotopic spanning)

=(𝒞):=inf{n(S):S𝒮(𝒞)}.\ell=\ell(\mathcal{C}):=\inf\big{\{}\mathcal{H}^{n}(S):S\in\mathcal{S}(\mathcal{C})\big{\}}\,. (1.7)

Existence of minimizers of \ell holds as soon as <\ell<\infty, and minimizers SS of \ell are Almgren minimal sets in Ω\Omega [HP16, DLGM17a] that are indeed going to exhibit Plateau-type singularities (this is easily seen in the plane, but see also [BM21] for a higher dimensional example). Moreover, given a same 𝐖\mathbf{W}, different choices of 𝒞\mathcal{C} are possible and can lead to different minimizers, see

Refer to caption
\begin{picture}(5148.0,1710.0)(2189.0,-1445.0)\end{picture}
Figure 1.3. The dashed lines denote the embeddings of SS1\SS^{1} whose homotopy classes relative to Ω\Omega generate different spanning classes 𝒞\mathcal{C}, to which there correspond different minimizers of \ell.

Figure 1.3. Finally, the approach is robust enough to provide the starting point for several important extensions [DPDRG16, DR18, HP17, FK18, DLDRG19, DPDRG20], including higher codimension, anisotropic energies, etc.

The study of soap films as minimizers of Gauss’s capillarity energy with small volume and under homotopic spanning conditions has been initiated in [MSS19, KMS22a], with the introduction of the model

ψ(v):=inf{n(ΩE):|E|=v,ΩE is 𝒞-spanning 𝐖},\psi(v):=\inf\Big{\{}\mathcal{H}^{n}(\Omega\cap\partial E):|E|=v\,,\,\,\mbox{$\Omega\cap\partial E$ is $\mathcal{C}$-spanning $\mathbf{W}$}\Big{\}}\,, (1.8)

where EΩE\subset\Omega is an open set with smooth boundary. Without the spanning condition, at small volumes, minimizers of n(ΩE)\mathcal{H}^{n}(\Omega\cap\partial E) would be small diffeomorphic images of half-balls [MM16]. However, the introduction of the 𝒞\mathcal{C}-spanning constraint rules out small droplets, and forces the exploration of a different part of the energy landscape of n(ΩE)\mathcal{H}^{n}(\Omega\cap\partial E). As informally discussed in Section 1.1, this leads to the emergence of generalized minimizers (K,E)(K,E). More precisely, in [KMS22a] the existence is proved of (K,E)(K,E) in the class

𝒦={(K,E):K is relatively closed and n-rectifiable in ΩE is open,\displaystyle\mathcal{K}=\Big{\{}(K,E):\mbox{$K$ is relatively closed and $\mathcal{H}^{n}$-rectifiable in $\Omega$, $E$ is open,} (1.9)
E has finite perimeter in Ω, and Ωcl(E)=ΩEK},\displaystyle\hskip 76.82234pt\mbox{$E$ has finite perimeter in $\Omega$, and $\Omega\cap\mathrm{cl}\,(\partial^{*}E)=\Omega\cap\partial E\subset K$}\Big{\}}\,,

(where E\partial^{*}E denotes the reduced boundary of EE) such that, for every competitor EE^{\prime} in ψ(v)\psi(v), it holds

n(ΩE)+2n(Ω(KE))n(ΩE).\mathcal{H}^{n}(\Omega\cap\partial^{*}E)+2\,\mathcal{H}^{n}(\Omega\cap(K\setminus\partial^{*}E))\leq\mathcal{H}^{n}(\Omega\cap\partial E^{\prime})\,. (1.10)

Starting from (1.10) one can apply Allard’s regularity theorem [All72] and various ad hoc comparison arguments [KMS21, KMS22b] to prove that ΩE\Omega\cap\partial^{*}E is a smooth hypersurface with constant mean curvature (negative if n(KE)>0\mathcal{H}^{n}(K\setminus\partial^{*}E)>0), Ω(EE)\Omega\cap(\partial E\setminus\partial^{*}E) has empty interior in KK, and that K(ΣE)K\setminus(\Sigma\cup\partial E) is a smooth minimal hypersurface, where Σ\Sigma is a closed set with codimension at least 88.

1.3. Measure theoretic homotopic spanning

In a nutshell, the idea behind our measure theoretic revision the Harrison–Pugh homotopic spanning condition is the following. Rather than asking that Sγ(SS1)S\cap\gamma(\SS^{1})\neq\varnothing for every γ𝒞\gamma\in\mathcal{C}, as done in Definition A, we shall replace γ\gamma with an open “tube” TT containing γ(SS1)\gamma(\SS^{1}), and ask that SS, with the help of a generic “slice” T[s]T[s] of TT, “disconnects” TT itself into two nontrivial regions T1T_{1} and T2T_{2}; see

Refer to caption
TTSSSST[s]T[s]γ(s)\gamma(s)γ\gamma𝐖\mathbf{W}(a)(b)
Figure 1.4. (a) Homotopic spanning according to Harrison–Pugh: SS must intersect every curve γ𝒞\gamma\in\mathcal{C}, in particular, the 𝒞\mathcal{C}-spanning property may be lost by removing a single point from SS; (b) Homotopic spanning based on essential connectedness: for a.e. section T[s]T[s] of the tube TT around a curve γ𝒞\gamma\in\mathcal{C}, the union T[s]ST[s]\cup S (essentially) disconnects TT (i.e., divides TT into two non-trivial parts, depicted here with two different shades of gray).

Figure 1.4. The key to make this idea work is, of course, giving a proper meaning to the word “disconnects”.

To this end, we recall the notion of essential connectedness introduced in [CCDPM17, CCDPM14] in the study of the rigidity of equality cases in Gaussian and Euclidean perimeter symmetrization inequalities. Essential connectedness is the “right” notion to deal with such problems since it leads to the formulation of sharp rigidity theorems, and can indeed be used to address other rigidity problems (see [CPS20, Per22, Dom23]). This said, it seems remarkable that the very same notion of what it means for “one Borel set to disconnect another Borel set” proves to be extremely effective also in the context of the present paper, which is of course very far from the context of symmetrization theory.

Denoting by T(t)T^{{\scriptscriptstyle{(t)}}} (0t10\leq t\leq 1) the points of density tt of a Borel set Tn+1T\subset\mathbb{R}^{n+1} (i.e., xT(t)x\in T^{{\scriptscriptstyle{(t)}}} if and only if |TBr(x)|/ωn+1rn+1t|T\cap B_{r}(x)|/\omega_{n+1}\,r^{n+1}\to t as r0+r\to 0^{+}, where ωk\omega_{k} is the Lebesgue measure of the unit ball in k\mathbb{R}^{k}), and by eT=n+1(T(0)T(1))\partial^{e}T=\mathbb{R}^{n+1}\setminus(T^{{\scriptscriptstyle{(0)}}}\cup T^{{\scriptscriptstyle{(1)}}}) the essential boundary of TT, given Borel sets SS, TT, T1T_{1} and T2T_{2} in n+1\mathbb{R}^{n+1}, and given n0n\geq 0, we say that SS essentially disconnects TT into {T1,T2}\{T_{1},T_{2}\}, if

{T1,T2} is a non-trivial Borel partition of T,and T(1)eT1eT2 is n-contained in S.\begin{split}&\mbox{$\{T_{1},T_{2}\}$ is a non-trivial Borel partition of $T$}\,,\\ &\mbox{and $T^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}$ is $\mathcal{H}^{n}$-contained in $S$}\,.\end{split} (1.11)

(For example, if KK is a set of full 1\mathcal{L}^{1}-measure in [1,1][-1,1], then S=K×{0}S=K\times\{0\} essentially disconnects the unit disk in 2\mathbb{R}^{2}.) Moreover, we say that TT is essentially connected333Whenever TT is of locally finite perimeter, being essentially connected is equivalent to being indecomposable. if \varnothing does not essentially disconnect TT. The requirement that {T1,T2}\{T_{1},T_{2}\} is a non-trivial Borel partition of TT means that |TΔ(T1T2)|=0|T\Delta(T_{1}\cup T_{2})|=0 and |T1||T2|>0|T_{1}|\,|T_{2}|>0. By saying that “EE is n\mathcal{H}^{n}-contained in FF” we mean that n(EF)=0\mathcal{H}^{n}(E\setminus F)=0. We also notice that, in (1.11), we have T(1)eT1eT2=T(1)eTiT^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}=T^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}T_{i} (i=1,2i=1,2), a fact that is tacitly and repeatedly considered in the use of (1.11) in order to shorten formulas.

With this terminology in mind, we introduce the following definition:

Definition B (Measure theoretic homotopic spanning).

Given a closed set 𝐖\mathbf{W} and a spanning class 𝒞\mathcal{C} for 𝐖\mathbf{W}, the tubular spanning class 𝒯(𝒞)\mathcal{T}(\mathcal{C}) associated to 𝒞\mathcal{C} is the family of triples (γ,Φ,T)(\gamma,\Phi,T) such that γ𝒞\gamma\in\mathcal{C}, T=Φ(𝕊1×B1n)T=\Phi(\mathbb{S}^{1}\times B_{1}^{n}), and444Here B1n={xn:|x|<1}B_{1}^{n}=\{x\in\mathbb{R}^{n}:|x|<1\} and SS1={s2:|s|=1}\SS^{1}=\{s\in\mathbb{R}^{2}:|s|=1\}.

Φ:𝕊1×clB1nΩ is a diffeomorphism with Φ|𝕊1×{0}=γ.\textup{$\Phi:\mathbb{S}^{1}\times\mathrm{cl}\,B_{1}^{n}\to\Omega$ is a diffeomorphism with ${\left.\kern-1.2pt\Phi\right|_{\mathbb{S}^{1}\times\{0\}}}=\gamma$}\,.

When (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}), the slice of TT defined by s𝕊1s\in\mathbb{S}^{1} is

T[s]=Φ({s}×B1n).T[s]=\Phi(\{s\}\times B_{1}^{n})\,.

Finally, we say that a Borel set SΩS\subset\Omega is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} if for each (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}), 1\mathcal{H}^{1}-a.e. s𝕊1s\in\mathbb{S}^{1} has the following property:

for n\mathcal{H}^{n}-a.e. xT[s]x\in T[s]
\exists a partition {T1,T2}\{T_{1},T_{2}\} of TT s.t. xeT1eT2x\in\partial^{e}T_{1}\cap\partial^{e}T_{2} (1.12)
and s.t. ST[s] essentially disconnects T into {T1,T2}.\displaystyle\mbox{and s.t. $S\cup T[s]$ essentially disconnects $T$ into $\{T_{1},T_{2}\}$}\,.

Before commenting on (1.12), we notice that the terminology of Definition B is coherent with that of Definition A thanks to the following theorem.

Theorem 1.1.

Given a closed set 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1}, a spanning class 𝒞\mathcal{C} for 𝐖\mathbf{W}, and a set SS relatively closed in Ω\Omega, then SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} in the sense of Definition A if and only if SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} in the sense of Definition B.

Theorem 1.1 is proved in Appendix A. There we also comment on the delicate reason why, in formulating (1.12), the partition {T1,T2}\{T_{1},T_{2}\} must be allowed to depend on specific points xT[s]x\in T[s]. This would not seem necessary by looking at the simple situation depicted in Figure 1.4, but it is actually so when dealing with more complex situations; see Figure A.1.

Homotopic spanning according to Definition B is clearly stable under modifications of SS by n\mathcal{H}^{n}-negligible sets, but there is more to it.

Refer to caption
(b)(b)(a)(a)UUSSU3U_{3}U2U_{2}U1U_{1}
Figure 1.5. An example of induced essential partition. The union of the boundaries of the UiU_{i}’s (inside of UU) is contained in SS, and the containment may be strict. However, the part of SS not contained in UiUiU\cap\bigcup_{i}\partial U_{i} is not such to disconnect any of the UiU_{i}’s. In particular, each UiU_{i} is essentially connected.

Indeed, even a notion like “n(ST)>0\mathcal{H}^{n}(S\cap T)>0 for every T𝒯(𝒞)T\in\mathcal{T}(\mathcal{C})” would be stable under modifications by n\mathcal{H}^{n}-negligible sets, and would probably look more appealing in its simplicity. The catch, of course, is finding an extension of Definition A for which compactness theorems, like Theorem 1.4 below, hold true. This is evidently not the case, for example, if one tries to work with a notion like “n(ST)>0\mathcal{H}^{n}(S\cap T)>0 for every T𝒯(𝒞)T\in\mathcal{T}(\mathcal{C})”.

The first key insight on Definition B is that, if restricted to Borel sets SS that are locally n\mathcal{H}^{n}-finite in Ω\Omega, then it can be reformulated in terms of partitions into indecomposable sets of finite perimeter. This is the content of the following theorem, whose case S=S=\varnothing corresponds to the standard decomposition theorem for sets of finite perimeter [ACMM01, Theorem 1]. For an illustration of this result, see Figure 1.5.

Theorem 1.2 (Induced essential partitions (Section 2)).

If Un+1U\subset\mathbb{R}^{n+1} is a bounded set of finite perimeter and Sn+1S\subset\mathbb{R}^{n+1} is a Borel set with n(SU(1))<\mathcal{H}^{n}(S\cap U^{\scriptscriptstyle{(1)}})<\infty, then there exists a unique555Uniqueness is meant modulo relabeling and modulo Lebesgue negligible modifications of the UiU_{i}’s. essential partition {Ui}i\{U_{i}\}_{i} of UU induced by SS, that is to say, {Ui}i\{U_{i}\}_{i} is a countable partition of UU modulo Lebesgue negligible sets such that, for each ii, SS does not essentially disconnect UiU_{i}.

Given UU and SS as in the statement of Theorem 1.2 we can define666Uniquely modulo n\mathcal{H}^{n}-null sets thanks to Federer’s theorem recalled in (1.37) below. the union of the (reduced) boundaries (relative to UU) of the essential partition induced by SS on UU by setting777Given a Borel set EE, we denote by E\partial^{*}E its reduced boundary relative to the maximal open set AA wherein EE has locally finite perimeter.

UBEP(S;U)=U(1)iUi.{\rm UBEP}(S;U)=U^{\scriptscriptstyle{(1)}}\cap\bigcup_{i}\partial^{*}U_{i}\,. (1.13)

Two properties of UBEP{\rm UBEP}’s which well illustrate the concept are: first, if (S)\mathcal{R}(S) denotes the rectifiable part of SS, then UBEP(S;U){\rm UBEP}(S;U) is n\mathcal{H}^{n}-equivalent to UBEP((S);U){\rm UBEP}(\mathcal{R}(S);U); second, if SS^{*} is n\mathcal{H}^{n}-contained in SS, then UBEP(S;U){\rm UBEP}(S;U) is n\mathcal{H}^{n}-contained in UBEP(S;U){\rm UBEP}(S;U); both properties are proved in Theorem 2.1 (an expanded restatement of Theorem 1.2).

We can use the concepts just introduced to provide an alternative and technically more workable characterization of homotopic spanning in the measure theoretic setting. This is the content of our first main result, which is illustrated in Figure 1.6.

Theorem 1.3 (Homotopic spanning for locally n\mathcal{H}^{n}-finite sets (Section 3)).

If 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} is a closed set in n+1\mathbb{R}^{n+1}, 𝒞\mathcal{C} is a spanning class for 𝐖\mathbf{W}, and SΩS\subset\Omega is locally n\mathcal{H}^{n}-finite in Ω\Omega, then SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} if and only if for every (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}) we have that, for 1\mathcal{H}^{1}-a.e. sSS1s\in\SS^{1},

T[s] is n-contained in UBEP(ST[s];T).\displaystyle\mbox{$T[s]$ is $\mathcal{H}^{n}$-contained in ${\rm UBEP}(S\cup T[s];T)$}\,. (1.14)
Refer to caption
TTSSSST[s]T[s]T[s]T[s]𝐖\mathbf{W}(b)(b)(a)(a)U1U_{1}U2U_{2}
Figure 1.6. With 𝐖\mathbf{W} consisting of two disks in the plane, and TT a test tube for the 𝒞\mathcal{C}-spanning condition: (a) SS consists of a segment with a gap: since the gap is inside of TT, the essential partition of TT induced by ST[s]S\cup T[s] consists of only one set, U1=TU_{1}=T, so that TU1=T\cap\partial^{*}U_{1}=\varnothing and (1.14) cannot hold; (b) SS consists of a full segment and in this case (with the possible exception of a choice of ss such that T[s]T[s] is contained in SS), the essential partition of TT induced by ST[s]S\cup T[s] consists of two sets {U1,U2}\{U_{1},U_{2}\}, such that T[s]TU1U2T[s]\subset T\cap\partial^{*}U_{1}\cap\partial^{*}U_{2}; in this case (1.14) holds.

1.4. Direct Method on generalized soap films and Gauss’ capillarity

The most convenient setting for addressing the existence of minimizers in Gauss’ capillarity theory is of course that of sets of finite perimeter [Fin86, Mag12]. However, if the notion of homotopic spanning is limited to closed sets, as it is the case when working with Definition A, then one cannot directly use homotopic spanning on sets of finite perimeter, and this is the reason behind the specific formulation (1.8) of ψ(v)\psi(v) used in [MSS19, KMS22a]. Equipped with Definition B we can now formulate Gauss’ capillarity theory with homotopic spanning conditions directly on sets of finite perimeter. We shall actually consider two different possible formulations

ψbk(v)=inf{n(ΩE):|E|=v and Ω(EE(1)) is 𝒞-spanning 𝐖},\displaystyle\psi_{\rm bk}(v)=\inf\Big{\{}\mathcal{H}^{n}(\Omega\cap\partial^{*}E):\mbox{$|E|=v$ and $\Omega\cap(\partial^{*}E\cup E^{\scriptscriptstyle{(1)}})$ is $\mathcal{C}$-spanning $\mathbf{W}$}\Big{\}}\,,
ψbd(v)=inf{n(ΩE):|E|=v and ΩE is 𝒞-spanning 𝐖},\displaystyle\psi_{\rm bd}(v)=\inf\Big{\{}\mathcal{H}^{n}(\Omega\cap\partial^{*}E):\mbox{$|E|=v$ and $\Omega\cap\partial^{*}E$ is $\mathcal{C}$-spanning $\mathbf{W}$}\Big{\}}\,,

where the subscripts “bk” and “bd” stand to indicate that the spanning is prescribed via the bulk of EE (that is, in measure theoretic terms, via the set Ω(EE(1))\Omega\cap(\partial^{*}E\cup E^{\scriptscriptstyle{(1)}}) or via the (reduced) boundary of EE. Inspired by the definition of the class 𝒦\mathcal{K} introduced in (1.9), we also introduce the class 𝒦B\mathcal{K}_{\rm B} of generalized soap films defined by

𝒦B={(K,E):K and E are Borel subsets of Ω,\displaystyle\mathcal{K}_{\rm B}=\Big{\{}(K,E):\mbox{$K$ and $E$ are Borel subsets of $\Omega$,} (1.15)
E has locally finite perimeter in Ω and EΩnK}.\displaystyle\hskip 76.82234pt\mbox{$E$ has locally finite perimeter in $\Omega$ and $\partial^{*}E\cap\Omega\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}K$}\Big{\}}\,.

Here the subscript “B” stands for “Borel”, and 𝒦B\mathcal{K}_{\rm B} stands as a sort of measure-theoretic version of 𝒦\mathcal{K}.

In the companion paper [Nov23] the following relaxation formulas for problems ψbk\psi_{\rm bk} and ψbd\psi_{\rm bd} are proved,

ψbk(v)=Ψbk(v),ψbd(v)=Ψbd(v),v>0,\psi_{\rm bk}(v)=\Psi_{\rm bk}(v)\,,\qquad\psi_{\rm bd}(v)=\Psi_{\rm bd}(v)\,,\qquad\forall v>0\,, (1.16)

where the following minimization problems on 𝒦B\mathcal{K}_{\rm B} are introduced

Ψbk(v)=inf{bk(K,E):(K,E)𝒦B,|E|=v,KE(1) is 𝒞-spanning 𝐖},\displaystyle\Psi_{\rm bk}(v)=\inf\Big{\{}\mathcal{F}_{\rm bk}(K,E):(K,E)\in\mathcal{K}_{\rm B}\,,|E|=v\,,\mbox{$K\cup E^{{\scriptscriptstyle{(1)}}}$ is $\mathcal{C}$-spanning $\mathbf{W}$}\Big{\}}\,,\hskip 28.45274pt (1.17)
Ψbd(v)=inf{bd(K,E):(K,E)𝒦B,|E|=v,K is 𝒞-spanning 𝐖}.\displaystyle\Psi_{\rm bd}(v)=\inf\Big{\{}\mathcal{F}_{\rm bd}(K,E):(K,E)\in\mathcal{K}_{\rm B}\,,|E|=v\,,\mbox{$K$ is $\mathcal{C}$-spanning $\mathbf{W}$}\Big{\}}\,. (1.18)

Here bk\mathcal{F}_{\rm bk} and bd\mathcal{F}_{\rm bd} are the relaxed energies defined for (K,E)𝒦B(K,E)\in\mathcal{K}_{\rm B} and AΩA\subset\Omega as

bk(K,E;A)=2n(AKE(0))+n(AE),\displaystyle\mathcal{F}_{\rm bk}(K,E;A)=2\,\mathcal{H}^{n}(A\cap K\cap E^{{\scriptscriptstyle{(0)}}})+\mathcal{H}^{n}(A\cap\partial^{*}E)\,, (1.19)
bd(K,E;A)=2n(AKE)+n(AE),\displaystyle\mathcal{F}_{\rm bd}(K,E;A)=2\,\mathcal{H}^{n}(A\cap K\setminus\partial^{*}E)+\mathcal{H}^{n}(A\cap\partial^{*}E)\,, (1.20)

(We also set, for brevity, bk(K,E):=bk(K,E;Ω)\mathcal{F}_{\rm bk}(K,E):=\mathcal{F}_{\rm bk}(K,E;\Omega) and bd(K,E):=bd(K,E;Ω)\mathcal{F}_{\rm bd}(K,E):=\mathcal{F}_{\rm bd}(K,E;\Omega).) We refer to these problems, respectively, as the “bulk-spanning” or “boundary-spanning” Gauss’ capillarity models. In this paper we shall directly work with these relaxed models. In particular, the validity of (1.16), although of definite conceptual importance, is not playing any formal role in our deductions.

A first remark concerning the advantage of working with the relaxed problems Ψbk\Psi_{\rm bk} and Ψbd\Psi_{\rm bd} rather than with their “classical” counterparts ψbk\psi_{\rm bk} and ψbd\psi_{\rm bd} is that while the latter two with v=0v=0 are trivial (sets with zero volume have zero distributional perimeter), the problems Ψbk(0)\Psi_{\rm bk}(0) and Ψbd(0)\Psi_{\rm bd}(0) are actually non-trivial, equal to each other, and amount to a measure-theoretic version of the Harrison–Pugh formulation of Plateau’s problem \ell introduced in (1.7): more precisely, if we set

B:=Ψbk(0)2=Ψbd(0)2=inf{n(S):S is a Borel set 𝒞-spanning 𝐖},\ell_{\rm B}:=\frac{\Psi_{\rm bk}(0)}{2}=\frac{\Psi_{\rm bd}(0)}{2}=\inf\Big{\{}\mathcal{H}^{n}(S):\mbox{$S$ is a Borel set $\mathcal{C}$-spanning $\mathbf{W}$}\Big{\}}\,, (1.21)

then, by Theorem 1.1, we evidently have B\ell_{\rm B}\leq\ell; and, as we shall prove in the course of our analysis, we actually have that =B\ell=\ell_{\rm B} as soon as <\ell<\infty.

Our second main result concerns the applicability of the Direct Method on the competition classes of Ψbk(v)\Psi_{\rm bk}(v) and Ψbd(v)\Psi_{\rm bd}(v).

Theorem 1.4 (Direct Method for generalized soap films (Sections 4 and 5)).

Let 𝐖\mathbf{W} be a closed set in n+1\mathbb{R}^{n+1}, 𝒞\mathcal{C} a spanning class for 𝐖\mathbf{W}, {(Kj,Ej)}j\{(K_{j},E_{j})\}_{j} be a sequence in 𝒦B\mathcal{K}_{\rm B} such that supjn(Kj)<\sup_{j}\mathcal{H}^{n}(K_{j})<\infty, and let a Borel set EE and Radon measures μbk\mu_{\rm bk} and μbd\mu_{\rm bd} in Ω\Omega be such that EjlocEE_{j}\stackrel{{\scriptstyle\scriptscriptstyle{{\rm loc}}}}{{\to}}E and

n (ΩEj)+2n ((Kj)Ej(0))μbk,\displaystyle\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E_{j})+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(0)}})\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}\mu_{\rm bk}\,,
n (ΩEj)+2n ((Kj)Ej)μbd,\displaystyle\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E_{j})+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\mathcal{R}(K_{j})\setminus\partial^{*}E_{j})\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}\mu_{\rm bd}\,,

as jj\to\infty. Then:

(i) Lower semicontinuity: the sets

Kbk\displaystyle K_{\rm bk}\!\! :=\displaystyle:= (ΩE){xΩE(0):θn(μbk)(x)2},\displaystyle\!\!\big{(}\Omega\cap\partial^{*}E\big{)}\cup\Big{\{}x\in\Omega\cap E^{\scriptscriptstyle{(0)}}:\theta^{n}_{*}(\mu_{\rm bk})(x)\geq 2\Big{\}}\,,
Kbd\displaystyle K_{\rm bd}\!\! :=\displaystyle:= (ΩE){xΩE:θn(μbd)(x)2},\displaystyle\!\!\big{(}\Omega\cap\partial^{*}E\big{)}\cup\Big{\{}x\in\Omega\setminus\partial^{*}E:\theta^{n}_{*}(\mu_{\rm bd})(x)\geq 2\Big{\}}\,,

are such that (Kbk,E),(Kbd,E)𝒦B(K_{\rm bk},E),(K_{\rm bd},E)\in\mathcal{K}_{\rm B} and

μbk\displaystyle\mu_{\rm bk}\!\! \displaystyle\geq n (ΩE)+2n (KbkE(0)),\displaystyle\!\!\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E)+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{\rm bk}\cap E^{\scriptscriptstyle{(0)}})\,,
μbd\displaystyle\mu_{\rm bd}\!\! \displaystyle\geq n (ΩE)+2n (KbdE),\displaystyle\!\!\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E)+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{\rm bd}\setminus\partial^{*}E)\,,

with

lim infjbk(Kj,Ej)bk(Kbk,E),lim infjbd(Kj,Ej)bd(Kbd,E).\liminf_{j\to\infty}\mathcal{F}_{\rm bk}(K_{j},E_{j})\geq\mathcal{F}_{\rm bk}(K_{\rm bk},E)\,,\qquad\liminf_{j\to\infty}\mathcal{F}_{\rm bd}(K_{j},E_{j})\geq\mathcal{F}_{\rm bd}(K_{\rm bd},E)\,.

(ii) Closure: we have that

if KjEj(1)K_{j}\cup E_{j}^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} for every jj,
then KbkE(1) is 𝒞-spanning 𝐖,\displaystyle\mbox{then $K_{\rm bk}\cup E^{\scriptscriptstyle{(1)}}$ is $\mathcal{C}$-spanning $\mathbf{W}$}\,,

and that

if KjK_{j} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} for every jj,
then Kbd is 𝒞-spanning 𝐖.\displaystyle\mbox{then $K_{\rm bd}$ is $\mathcal{C}$-spanning $\mathbf{W}$}\,.

The delicate part of Theorem 1.4 is proving the closure statements. This will require first to extend the characterization of homotopic spanning from locally n\mathcal{H}^{n}-finite sets to generalized soap films (Theorem 3.1), and then to discuss the behavior under weak-star convergence of the associated Radon measures of the objects appearing in conditions like (1.14) (Theorem 4.1).

1.5. Existence of minimizers in Ψbk(v)\Psi_{\rm bk}(v) and convergence to \ell

From this point onward, we focus our analysis on the bulk-spanning relaxation Ψbk(v)\Psi_{\rm bk}(v) of Gauss’ capillarity. There are a few important reasons for this choice: (i) from the point of view of physical modeling, working with the boundary or with the bulk spanning conditions seem comparable; (ii) the fact that Ψbk(0)=Ψbd(0)\Psi_{\rm bk}(0)=\Psi_{\rm bd}(0) suggest that, at small values of vv, the two problems should actually be equivalent (have the same infima and the same minimizers); (iii) the bulk spanning variant is the one which is relevant for the approximation of Plateau-type singularities with solutions of the Allen–Cahn equations discussed in [MNR23a]; (iv) despite their similarities, carrying over the following theorems for both problems would require the repeated introduction of two versions of many arguments, with a significant increase in length, and possibly with at the expense of clarity.

The following theorem provides the starting point in the study of Ψbk(v)\Psi_{\rm bk}(v).

Theorem 1.5 (Existence of minimizers and vanishing volume limit for Ψbk\Psi_{\rm bk} (Section 6)).

If 𝐖\mathbf{W} is a compact set in n+1\mathbb{R}^{n+1} and 𝒞\mathcal{C} is a spanning class for 𝐖\mathbf{W} such that <\ell<\infty, then

B=,\ell_{\rm B}=\ell\,, (1.22)

and, moreover:

(i) Existence of minimizers and Euler–Lagrange equation: for every v>0v>0 there exist minimizers (K,E)(K,E) of Ψbk(v)\Psi_{\rm bk}(v) such that (K,E)𝒦(K,E)\in\mathcal{K} and both EE and KK are bounded; moreover, there is λ\lambda\in\mathbb{R} such that

λEXνE𝑑n=EdivKX𝑑n+2KE(0)divKX𝑑n,\displaystyle\lambda\int_{\partial^{*}E}X\cdot\nu_{E}\,d\mathcal{H}^{n}=\int_{\partial^{*}E}\mathrm{div}^{K}\,X\,d\mathcal{H}^{n}+2\int_{K\cap E^{{\scriptscriptstyle{(0)}}}}\mathrm{div}^{K}\,X\,d\mathcal{H}^{n}\,, (1.23)

for every XCc1(n+1;n+1)X\in C^{1}_{c}(\mathbb{R}^{n+1};\mathbb{R}^{n+1}) with XνΩ=0X\cdot\nu_{\Omega}=0 on Ω\partial\Omega;

(ii) Regularity from the Euler–Lagrange equations: if (K,E)𝒦(K,E)\in\mathcal{K} is a minimizer of either Ψbk(v)\Psi_{\rm bk}(v), then there is a closed set ΣK\Sigma\subset K, with empty interior in KK, such that KΣK\setminus\Sigma is a smooth hypersurface; moreover, K(ΣE)K\setminus(\Sigma\cup\partial E) is a smooth minimal hypersurface, ΩE\Omega\cap\partial^{*}E is a smooth hypersurface with mean curvature constantly equal to λ\lambda, and n(ΣE)=0\mathcal{H}^{n}(\Sigma\setminus\partial E)=0; in particular, Ω(EE)\Omega\cap(\partial E\setminus\partial^{*}E) has empty interior in KK;

(iii) Convergence to the Plateau problem: if (Kj,Ej)(K_{j},E_{j}) is a sequence of minimizers for Ψbk(vj)\Psi_{\rm bk}(v_{j}) with vj0+v_{j}\to 0^{+}, then there exists a minimizer SS of \ell such that, up to extracting subsequences, as Radon measures in Ω\Omega,

n (EjΩ)+2n (KjEj(0))2n S,\displaystyle\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}E_{j}\cap\Omega)+2\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{j}\cap E_{j}^{{\scriptscriptstyle{(0)}}})\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}2\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S\,, (1.24)

as jj\to\infty; In particular, Ψbk(v)2=Ψbk(0)\Psi_{\rm bk}(v)\to 2\,\ell=\Psi_{\rm bk}(0) as v0+v\to 0^{+}.

The conclusions of Theorem 1.5 about Ψbk(v)\Psi_{\rm bk}(v) can be read in parallel to the conclusions about ψ(v)\psi(v) obtained in [KMS22a]. The crucial difference is that, in place of the “weak” minimality inequality (1.10), which in this context would be equivalent to bk(K,E)n(ΩE)\mathcal{F}_{\rm bk}(K,E)\leq\mathcal{H}^{n}(\Omega\cap\partial^{*}E^{\prime}) for every competitor EE^{\prime} in ψbk(v)\psi_{\rm bk}(v), we now have the proper minimality inequality

bk(K,E)bk(K,E)\mathcal{F}_{\rm bk}(K,E)\leq\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime}) (1.25)

for every competitor (K,E)(K^{\prime},E^{\prime}) in Ψbk(v)\Psi_{\rm bk}(v). Not only the final conclusion is stronger, but the proof is also entirely different: whereas [KMS22a] required the combination of a whole bestiary of specific competitors (like the cup, cone, and slab competitors described therein) with the full force of Preiss’ theorem, the approach presented here seems more robust as it does not exploit any specific geometry, and it is squarely rooted in the basic theory of sets of finite perimeter.

1.6. Equilibrium across transition lines in wet soap films

We now formalize the validation of (1.1) for soap films in the form of a sharp regularity theorem for minimizers (K,E)(K,E) of Ψbk(v)\Psi_{\rm bk}(v).

The starting point to obtain this result is the connection between homotopic spanning and partitions into indecomposable sets of finite perimeter established in Theorem 1.3/Theorem 3.1. This connection hints at the possibility of showing that if (K,E)(K,E) is a minimizer of Ψbk(v)\Psi_{\rm bk}(v), then the elements {Ui}i\{U_{i}\}_{i} of the essential partition of Ω\Omega induced by KE(1)K\cup E^{\scriptscriptstyle{(1)}} are actually (Λ,r0)(\Lambda,r_{0})-minimizers of the perimeter in Ω\Omega, i.e., there exist Λ\Lambda and r0r_{0} positive constants such that

P(Ui;Br(x))P(V;Br(x))+Λ|VΔUi|,P(U_{i};B_{r}(x))\leq P(V;B_{r}(x))+\Lambda\,|V\Delta U_{i}|\,,

whenever VΔUiΩV\Delta U_{i}\subset\!\subset\Omega and diam(VΔUi)<r0{\rm diam}\,(V\Delta U_{i})<r_{0}. The reason why this property is not obvious is that proving the (Λ,r0)(\Lambda,r_{0})-minimality of UiU_{i} requires working with arbitrary local competitors ViV_{i} of UiU_{i}. However, when working with homotopic spanning conditions, checking the admissibility of competitors is the notoriously delicate heart of the matter – as reflected in the fact that only very special classes of competitors have been considered in the literature (see, e.g., the cup and cone competitors and the Lipschitz deformations considered in [DLGM17a], the slab competitors and exterior cup competitors of [KMS22a], etc.).

The idea used to overcome this difficulty, which is illustrated in

Refer to caption
ViV_{i}UiU_{i}(Vi)(Ui)(\partial^{*}V_{i})\setminus(\partial^{*}U_{i})EE^{\prime}VjV_{j}VkV_{k}Br(x)B_{r}(x)UjU_{j}UkU_{k}EE
Figure 1.7. On the left, a minimizer (K,E)(K,E) of Ψbk(v)\Psi_{\rm bk}(v), and the essential partition induced by (K,E)(K,E) in a ball Br(x)B_{r}(x); the multiplicity 22 part of KBr(x)K\cap B_{r}(x) are depicted with bold lines, to distinguish them from the multiplicity one parts in Br(x)EB_{r}(x)\cap\partial^{*}E. On the right, a choice of (K,E)(K^{\prime},E^{\prime}) that guarantees both the energy gap identity (1.26) and the n\mathcal{H}^{n}-containment (1.27) needed to preserve homotopic spanning. The volume constraint can of course be restored as a lower order perimeter perturbation by taking a diffeomorphic image of (K,E)(K^{\prime},E^{\prime}), an operation that trivially preserves homotopic spanning.

Figure 1.7, is the following. By Theorem 1.2, we can locally represent bk(K,E;Br(x))\mathcal{F}_{\rm bk}(K,E;B_{r}(x)) as the sum of perimeters P(Ui;Br(x))+P(Uj;Br(x))+P(Uk;Br(x))P(U_{i};B_{r}(x))+P(U_{j};B_{r}(x))+P(U_{k};B_{r}(x)). Given a local competitor ViV_{i} for UiU_{i} we can carefully define a competitor (K,E)(K^{\prime},E^{\prime}) so that the elements of the essential partition induced by K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}} in Ω\Omega, that can be used to represent bk(K,E;Br(x))\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime};B_{r}(x)) as the sum P(Vi;Br(x))+P(Vj;Br(x))+P(Vk;Br(x))P(V_{i};B_{r}(x))+P(V_{j};B_{r}(x))+P(V_{k};B_{r}(x)), are such that

bk(K,E;Br(x))bk(K,E;Br(x))=P(V;Br(x))P(Ui;Br(x)).\displaystyle\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime};B_{r}(x))-\mathcal{F}_{\rm bk}(K,E;B_{r}(x))=P(V;B_{r}(x))-P(U_{i};B_{r}(x))\,. (1.26)

The trick is that by suitably defining KK^{\prime} and EE^{\prime} we can recover the entirety of Br(x)UjB_{r}(x)\cap\partial^{*}U_{j} and Br(x)UkB_{r}(x)\cap\partial^{*}U_{k} by attributing different parts of these boundaries to different terms in the representation of bk(K,E;Br(x))\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime};B_{r}(x)). In other words we are claiming that things can be arranged so that we still have

Br(x)(UjUk)nK(E)(1).B_{r}(x)\cap\big{(}\partial^{*}U_{j}\cap\partial^{*}U_{k})\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}}\,. (1.27)

The fact that we have been able to preserve all but one reduced boundary among those of the elements of the essential partition of Br(x)B_{r}(x) induced by (K,E)(K,E) is enough to shows that K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}} is still 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} by means of Theorem 1.3/Theorem 3.1.

By the regularity theory of (Λ,r0)(\Lambda,r_{0})-perimeter minimizers (see, e.g. [Mag12, Part III]) we can deduce the C1,αC^{1,\alpha}-regularity of the elements of the partition (away from a closed singular set with area minimizing dimensional bounds). This is already sufficient to prove the continuity of the normal across Ω(EE)\Omega\cap(\partial E\setminus\partial^{*}E), but it also allows us to invoke the regularity theory for free boundaries in the double membrane problem, and to obtain the following sharp regularity result, with which we conclude our introduction.

Theorem 1.6 (Equilibrium along transition lines for soap films (Section 7)).

If 𝐖\mathbf{W} is a compact set in n+1\mathbb{R}^{n+1}, 𝒞\mathcal{C} is a spanning class for 𝐖\mathbf{W} such that <\ell<\infty, v>0v>0, and (K,E)(K_{*},E_{*}) is a minimizer of Ψbk(v)\Psi_{\rm bk}(v), then there is (K,E)𝒦(K,E)\in\mathcal{K} such that KK is n\mathcal{H}^{n}-equivalent to KK_{*}, EE is Lebesgue equivalent to EE_{*}, (K,E)(K,E) is a minimizer of Ψbk(v)\Psi_{\rm bk}(v), both EE and KK are bounded, KEK\cup E is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, and

KE(1)=;K\cap E^{\scriptscriptstyle{(1)}}=\varnothing\,; (1.28)

in particular, KK is the disjoint union of ΩE\Omega\cap\partial^{*}E, Ω(EE)\Omega\cap(\partial E\setminus\partial^{*}E), and KEK\setminus\partial E.

Moreover, there is a closed set ΣK\Sigma\subset K with the following properties:

(i): Σ=\Sigma=\varnothing if 1n61\leq n\leq 6, Σ\Sigma is locally finite in Ω\Omega if n=7n=7, and s(Σ)=0\mathcal{H}^{s}(\Sigma)=0 for every s>n7s>n-7 if n8n\geq 8;

(ii): (ΩE)Σ(\Omega\cap\partial^{*}E)\setminus\Sigma is a smooth hypersurface with constant mean curvature (denoted by λ\lambda if computed with respect to νE\nu_{E});

(iii): (KE)Σ(K\setminus\partial E)\setminus\Sigma is a smooth minimal hypersurface;

(iv): if Ω(EE)Σ\Omega\cap(\partial E\setminus\partial^{*}E)\setminus\Sigma\neq\varnothing, then λ<0\lambda<0; moreover, for every xΩ(EE)Σx\in\Omega\cap(\partial E\setminus\partial^{*}E)\setminus\Sigma, KK is the union of two C1,1C^{1,1}-hypersurfaces that detach tangentially at xx; more precisely, there are r>0r>0, νSSn\nu\in\SS^{n}, u1,u2C1,1(𝐃rν(x))u_{1},u_{2}\in C^{1,1}(\mathbf{D}_{r}^{\nu}(x)) such that

u1(x)=u2(x)=0,u1u2 on 𝐃rν(x),u_{1}(x)=u_{2}(x)=0\,,\qquad\mbox{$u_{1}\leq u_{2}$ on $\mathbf{D}_{r}^{\nu}(x)$}\,,

with {u1<u2}\{u_{1}<u_{2}\} and int{u1=u2}{\rm int}\{u_{1}=u_{2}\} both non-empty, and

𝐂rν(x)K\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap K =\displaystyle= i=1,2{y+ui(y)ν:y𝐃rν(x)},\displaystyle\cup_{i=1,2}\big{\{}y+u_{i}(y)\,\nu:y\in\mathbf{D}_{r}^{\nu}(x)\big{\}}\,, (1.29)
𝐂rν(x)E\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap\partial^{*}E =\displaystyle= i=1,2{y+ui(y)ν:y{u1<u2}},\displaystyle\cup_{i=1,2}\big{\{}y+u_{i}(y)\nu:y\in\{u_{1}<u_{2}\}\big{\}}\,, (1.30)
𝐂rν(x)E\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap E =\displaystyle= {y+tν:t(u1(y),u2(y))}.\displaystyle\big{\{}y+t\,\nu:t\in\big{(}u_{1}(y),u_{2}(y)\big{)}\big{\}}\,. (1.31)

Here,

𝐃νr(x)=x+{yν:|y|<r},\displaystyle\mathbf{D}_{\nu}^{r}(x)=x+\{y\in\nu^{\perp}:|y|<r\}\,,
𝐂νr(x)=x+{y+tν:yν,|y|<r,|t|<r}.\displaystyle\mathbf{C}_{\nu}^{r}(x)=x+\{y+t\,\nu:y\in\nu^{\perp}\,,|y|<r\,,|t|<r\}\,.

(v): we have

Γ:=Ω(EE)=ΓregΓsing,ΓregΓsing=,\Gamma:=\Omega\cap(\partial E\setminus\partial^{*}E)=\Gamma_{\rm reg}\cup\Gamma_{\rm sing}\,,\qquad\Gamma_{\rm reg}\cap\Gamma_{\rm sing}=\varnothing\,,

where: Γreg\Gamma_{\rm reg} is relatively open in Γ\Gamma and for every xΓregx\in\Gamma_{\rm reg} there are r>0r>0 and β(0,1)\beta\in(0,1) such that ΓregBr(x)\Gamma_{\rm reg}\cap B_{r}(x) is a C1,βC^{1,\beta}-embedded (n1)(n-1)-dimensional manifold; Γsing\Gamma_{\rm sing} is relatively closed in Γ\Gamma and can be partitioned into a family {Γsingk}k=0n1\{\Gamma_{\rm sing}^{k}\}_{k=0}^{n-1} where, for each kk, Γsingk\Gamma_{\rm sing}^{k} is locally k\mathcal{H}^{k}-rectifiable in Ω\Omega.

1.7. Equilibrium across transition lines in wet foams

Based on the descriptions provided in [WH99, CCAE+13], an effective mathematical model for dry foams at equilibrium in a container is that of locally perimeter minimizing clusters, originating with different terminology in [Alm76], and presented in [Mag12, Part IV] as follows. Given an open set Ωn+1\Omega\subset\mathbb{R}^{n+1}, a locally perimeter minimizing clusters is a finite Lebesgue partition {Ui}i\{U_{i}\}_{i} of Ω\Omega into sets of finite perimeter such that, for some r0>0r_{0}>0,

iP(Ui;B)iP(Vi;B)\sum_{i}P(U_{i};B)\leq\sum_{i}P(V_{i};B) (1.32)

whenever BΩB\subset\!\subset\Omega is a ball with radius less than r0r_{0}, and {Vi}i\{V_{i}\}_{i} is a Lebesgue partition of Ω\Omega with ViΔUiBV_{i}\Delta U_{i}\subset\!\subset B and |Vi|=|Ui||V_{i}|=|U_{i}| for every ii. The previously cited results of Almgren and Taylor [Alm76, Tay76] imply that, up to modification of the UiU_{i}’s by sets of zero Lebesgue measure, when n=2n=2, K=ΩiUiK=\Omega\cap\bigcup_{i}\partial U_{i} is a closed subset of Ω\Omega that is locally C1,αC^{1,\alpha}-diffeomorphic to a plane, a YY-cone, or a TT-cone; moreover, the part of KK that is a surface is actually smooth and each of its connected component has constant mean curvature. Similar results holds when n=1n=1 (by elementary methods) and when n3n\geq 3 (by exploiting [CES22]).

The theory for the relaxed capillarity energy bk\mathcal{F}_{\rm bk} developed in this paper provides an option for modeling wet foams. Again based on the descriptions provided in [WH99, CCAE+13], the following seems to be a reasonable model for wet foams at equilibrium in a container. Given an open set Ωn+1\Omega\subset\mathbb{R}^{n+1} we model wet foams by introducing the class

𝒦foam\mathcal{K}_{{\rm foam}}

of those (K,E)𝒦B(K,E)\in\mathcal{K}_{\rm B} such that, for some positive constants Λ0\Lambda_{0} and r0r_{0},

bk(K,E;B)bk(K,E;B)+Λ0|EΔE|\mathcal{F}_{\rm bk}(K,E;B)\leq\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime};B)+\,\Lambda_{0}\,|E\Delta E^{\prime}| (1.33)

whenever BB is a ball compactly contained in Ω\Omega and with radius less than r0r_{0}, and (K,E)𝒦B(K^{\prime},E^{\prime})\in\mathcal{K}_{\rm B} is such that (KΔK)(EΔE)B(K\Delta K^{\prime})\cup(E\Delta E^{\prime})\subset\!\subset B and there are finite Lebesgue partitions {Ui}i\{U_{i}\}_{i} and {Ui}i\{U^{\prime}_{i}\}_{i} of BB induced, respectively, by KE(1)K\cup E^{\scriptscriptstyle{(1)}} and by K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}}, such that |Ui|=|Ui||U_{i}|=|U_{i}^{\prime}| for every ii. Notice that inclusion of the term Λ0|EΔE|\Lambda_{0}\,|E\Delta E^{\prime}| in (1.33) allows for the inclusion of energy perturbations due to gravity or other forces. Lemma 7.1 will clarify that by taking (K,E)𝒦foam(K,E)\in\mathcal{K}_{\rm foam} with |E|=0|E|=0 we obtain a slightly more general notion of dry foam than the one proposed in (1.32).

Theorem 1.7 (Equilibrium along transition lines for soap films (Section 8)).

If Ωn+1\Omega\subset\mathbb{R}^{n+1} is open and (K,E)𝒦foam(K_{*},E_{*})\in\mathcal{K}_{\rm foam}, then there is (K,E)𝒦𝒦foam(K,E)\in\mathcal{K}\cap\mathcal{K}_{\rm foam} such that KK is n\mathcal{H}^{n}-equivalent to KK_{*}, EE Lebesgue equivalent to EE_{*}, KE(1)=K\cap E^{\scriptscriptstyle{(1)}}=\varnothing, and such that, for every ball BΩB\subset\!\subset\Omega, the open connected components {Ui}i\{U_{i}\}_{i} of B(KE)B\setminus(K\cup E) are such that each UiU_{i} is (Lebesgue equivalent to an) open set with C1,αC^{1,\alpha}-boundary in BΣB\setminus\Sigma. Here Σ\Sigma is a closed subset of Ω\Omega with Σ=\Sigma=\varnothing if 1n61\leq n\leq 6, Σ\Sigma locally finite in Ω\Omega if n=7n=7, and s(Σ)=0\mathcal{H}^{s}(\Sigma)=0 for every s>n7s>n-7 if n8n\geq 8.

Organization of the paper

The sections of the paper contain the proofs of the main theorems listed above, as already specified in the statements. To these section we add three appendices. In Appendix A, as already noted, we prove the equivalence of Definition A and Definition B. In Appendix B we prove that, with some regularity of Ω\partial\Omega, every minimizing sequence of Ψbk(v)\Psi_{\rm bk}(v) is converging to a minimizers, without need for modifications at infinity: this is, strictly speaking, not needed to prove Theorem 1.5, but it is a result of its own conceptual interest, it will be crucial for the analysis presented in [MNR23a], and it is easily discussed here in light of the proof of Theorem 1.5. Finally, Appendix C contains an elementary lemma concerning the use of homotopic spanning in the plane that, to our knowledge, has not been proved in two dimensions.

Acknowledgements

We thank Guido De Philippis, Darren King, Felix Otto, Antonello Scardicchio, Salvatore Stuvard, and Bozhidar Velichkov for several interesting discussions concerning these problems. FM has been supported by NSF Grant DMS-2247544. FM, MN, and DR have been supported by NSF Grant DMS-2000034 and NSF FRG Grant DMS-1854344. MN has been supported by NSF RTG Grant DMS-1840314.

Notation

Sets and measures: We denote by Br(x)B_{r}(x) (resp., Brk(x)B_{r}^{k}(x)) the open ball of center xx and radius rr in n+1\mathbb{R}^{n+1} (resp., k\mathbb{R}^{k}), and omit (x)(x) when x=0x=0. We denote by cl(X)\mathrm{cl}\,(X), int(X){\rm int}(X), and Ir(X)I_{r}(X) the closure, interior and open ε\varepsilon-neighborhood of XkX\subset\mathbb{R}^{k}. We denote by n+1\mathcal{L}^{n+1} and s\mathcal{H}^{s} the Lebesgue measure and the ss-dimensional Hausdorff measure on n+1\mathbb{R}^{n+1}, s[0,n+1]s\in[0,n+1]. If EkE\subset\mathbb{R}^{k}, we set |E|=k(E)|E|=\mathcal{L}^{k}(E) and ωk=|B1k|\omega_{k}=|B_{1}^{k}|. We denote by E(t)E^{{\scriptscriptstyle{(t)}}}, t[0,1]t\in[0,1], the points of density tt of a Borel set En+1E\subset\mathbb{R}^{n+1}, so that EE is n+1\mathcal{L}^{n+1}-equivalent to E(1)E^{{\scriptscriptstyle{(1)}}}, and, for every pair of Borel sets E,Fn+1E,F\subset\mathbb{R}^{n+1},

(EF)(0)=E(0)F(0).(E\cup F)^{\scriptscriptstyle{(0)}}=E^{\scriptscriptstyle{(0)}}\cap F^{\scriptscriptstyle{(0)}}\,. (1.34)

We define by eE=n+1(E(0)E(1))\partial^{e}E=\mathbb{R}^{n+1}\setminus(E^{{\scriptscriptstyle{(0)}}}\cup E^{{\scriptscriptstyle{(1)}}}) the essential boundary of EE. Given Borel sets Ej,EΩE_{j},E\subset\Omega we write

EjE,EjlocE,E_{j}\to E\,,\qquad E_{j}\stackrel{{\scriptstyle\scriptscriptstyle{{\rm loc}}}}{{\to}}E\,,

when, respectively, |EjΔE|0|E_{j}\Delta E|\to 0 or |(EjΔE)Ω|0|(E_{j}\Delta E)\cap\Omega^{\prime}|\to 0 for every ΩΩ\Omega^{\prime}\subset\!\subset\Omega, as jj\to\infty. Given a Radon measure μ\mu on n+1\mathbb{R}^{n+1}, the kk-dimensional lower density of μ\mu is the Borel function θk(μ):n+1[0,]\theta^{k}_{*}(\mu):\mathbb{R}^{n+1}\to[0,\infty] defined by

θk(μ)(x)=lim infr0+μ(cl(Br(x)))ωkrk.\displaystyle\theta^{k}_{*}(\mu)(x)=\liminf_{r\to 0^{+}}\frac{\mu(\mathrm{cl}\,(B_{r}(x)))}{\omega_{k}r^{k}}\,.

We repeatedly use the fact that, if θk(μ)λ\theta^{k}_{*}(\mu)\geq\lambda on some Borel set KK and for some λ0\lambda\geq 0, then μλkK\mu\geq\lambda\,\mathcal{H}^{k}\llcorner K; see, e.g. [Mag12, Theorem 6.4].

Rectifiable sets: Given an integer 0kn+10\leq k\leq n+1, a Borel set Sn+1S\subset\mathbb{R}^{n+1} is locally k\mathcal{H}^{k}-rectifiable in an open set Ω\Omega if SS is locally k\mathcal{H}^{k}-finite in Ω\Omega and SS can be covered, modulo k\mathcal{H}^{k}-null sets, by a countable union of Lipschitz images of k\mathbb{R}^{k} in n+1\mathbb{R}^{n+1}. We say that SS is purely k\mathcal{H}^{k}-unrectifiable if k(SM)=0\mathcal{H}^{k}(S\cap M)=0 whenever MM is a Lipschitz image of k\mathbb{R}^{k} into n+1\mathbb{R}^{n+1}. Finally, we recall that if SS is a locally k\mathcal{H}^{k}-finite set in Ω\Omega, then there is a pair ((S),𝒫(S))(\mathcal{R}(S),\mathcal{P}(S)) of Borel sets, uniquely determined modulo k\mathcal{H}^{k}-null sets, and that are thus called, with a slight abuse of language, the rectifiable part and the unrectifiable part of SS, so that (S)\mathcal{R}(S) is locally k\mathcal{H}^{k}-rectifiable in Ω\Omega, 𝒫(S)\mathcal{P}(S) is purely k\mathcal{H}^{k}-unrectifiable, and S=(S)𝒫(S)S=\mathcal{R}(S)\cup\mathcal{P}(S); see, e.g. [Sim83, 13.1].

Sets of finite perimeter: If EE is a Borel set in n+1\mathbb{R}^{n+1} and D1ED1_{E} is the distributional derivative of the characteristic function of EE, then we set μE=D1E\mu_{E}=-D1_{E}. If AA is the largest open set of n+1\mathbb{R}^{n+1} such that μE\mu_{E} is a Radon measure in AA (of course it could be A=A=\varnothing), then EE is of locally finite perimeter in AA and the reduced boundary E\partial^{*}E of EE is defined as the set of those xAsptμEx\in A\cap{\rm spt}\mu_{E} such that μE(Br(x))/|μE|(Br(x))\mu_{E}(B_{r}(x))/|\mu_{E}|(B_{r}(x)) has a limit νE(x)SSn\nu_{E}(x)\in\SS^{n} as r0+r\to 0^{+}. Moreover, we have the general identity (see [Mag12, (12.12) &\& pag. 168])

Acl(E)=AsptμE={xA:0<|EBr(x)|<|Br(x)|r>0}AE.A\cap\mathrm{cl}\,(\partial^{*}E)=A\cap{\rm spt}\mu_{E}=\Big{\{}x\in A:0<|E\cap B_{r}(x)|<|B_{r}(x)|\,\,\forall r>0\Big{\}}\subset A\cap\partial E\,. (1.35)

By De Giorgi’s rectifiability theorem, E\partial^{*}E is locally n\mathcal{H}^{n}-rectifiable in AA, μE=νEn  (AE)\mu_{E}=\nu_{E}\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(A\cap\partial^{*}E) on AA, and EAE(1/2)AeE\partial^{*}E\subset A\cap E^{\scriptscriptstyle{(1/2)}}\subset A\cap\partial^{e}E, and

(Ex)/rlocHE,x:={yn+1:yνE(x)<0},as r0+.(E-x)/r\stackrel{{\scriptstyle\scriptscriptstyle{{\rm loc}}}}{{\to}}H_{E,x}:=\big{\{}y\in\mathbb{R}^{n+1}:y\cdot\nu_{E}(x)<0\big{\}}\,,\qquad\mbox{as $r\to 0^{+}$}\,. (1.36)

By a result of Federer,

A is n-contained in E(0)E(1)E;\mbox{$A$ is $\mathcal{H}^{n}$-contained in $E^{{\scriptscriptstyle{(0)}}}\cup E^{{\scriptscriptstyle{(1)}}}\cup\partial^{*}E$}\,; (1.37)

in particular, E\partial^{*}E is n\mathcal{H}^{n}-equivalent to AeEA\cap\partial^{e}E, a fact frequently used in the following. By Federer’s criterion for finite perimeter, if Ω\Omega is open and EE is a Borel set, then

n(ΩeE)<E is of finite perimeter in Ω,\mathcal{H}^{n}(\Omega\cap\partial^{e}E)<\infty\qquad\Rightarrow\qquad\mbox{$E$ is of finite perimeter in $\Omega$}\,, (1.38)

see [Fed69, 4.5.11]. If EE and FF are of locally finite perimeter in Ω\Omega open, then so are EFE\cup F, EFE\cap F, and EFE\setminus F, and by [Mag12, Theorem 16.3], we have

Ω(EF)=nΩ{(E(0)F)(F(0)E){νE=νF}},\displaystyle\Omega\cap\partial^{*}(E\cup F)\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\Omega\cap\Big{\{}\big{(}E^{\scriptscriptstyle{(0)}}\cap\partial^{*}F\big{)}\cup\big{(}F^{\scriptscriptstyle{(0)}}\cap\partial^{*}E\big{)}\cup\{\nu_{E}=\nu_{F}\}\Big{\}}\,, (1.39)
Ω(EF)=nΩ{(E(1)F)(F(1)E){νE=νF}},\displaystyle\Omega\cap\partial^{*}(E\cap F)\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\Omega\cap\Big{\{}\big{(}E^{\scriptscriptstyle{(1)}}\cap\partial^{*}F\big{)}\cup\big{(}F^{\scriptscriptstyle{(1)}}\cap\partial^{*}E\big{)}\cup\{\nu_{E}=\nu_{F}\}\Big{\}}\,, (1.40)
Ω(EF)=nΩ{(E(1)F)(F(0)E){νE=νF}},\displaystyle\Omega\cap\partial^{*}(E\setminus F)\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\Omega\cap\Big{\{}\big{(}E^{\scriptscriptstyle{(1)}}\cap\partial^{*}F\big{)}\cup\big{(}F^{\scriptscriptstyle{(0)}}\cap\partial^{*}E\big{)}\cup\{\nu_{E}=-\nu_{F}\}\Big{\}}\,, (1.41)

where {νE=±νF}:={xEF:νE(x)=±νF(x)}\{\nu_{E}=\pm\nu_{F}\}:=\{x\in\partial^{*}E\cap\partial^{*}F:\nu_{E}(x)=\pm\nu_{F}(x)\}. By exploiting Federer’s theorem (1.37), (1.39), (1.40), and (1.41) we can also deduce (the details are left to the reader)

(EF)(0)\displaystyle(E\cap F)^{\scriptscriptstyle{(0)}} =n\displaystyle\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=} E(0)F(0){νE=νF},\displaystyle E^{\scriptscriptstyle{(0)}}\cup F^{\scriptscriptstyle{(0)}}\cup\{\nu_{E}=-\nu_{F}\}\,, (1.42)
(EF)(0)\displaystyle(E\setminus F)^{\scriptscriptstyle{(0)}} =n\displaystyle\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=} E(0)F(1){νE=νF}.\displaystyle E^{\scriptscriptstyle{(0)}}\cup F^{\scriptscriptstyle{(1)}}\cup\{\nu_{E}=\nu_{F}\}\,. (1.43)

Finally, combining (1.39), (1.41), and (1.43), we find

(EΔF)=n(E)Δ(F).\partial^{*}(E\Delta F)\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}(\partial^{*}E)\Delta(\partial^{*}F)\,. (1.44)

Partitions: Given a Radon measure μ\mu on n+1\mathbb{R}^{n+1} and Borel set Un+1U\subset\mathbb{R}^{n+1} we say that {Ui}i\{U_{i}\}_{i} is a μ\mu-partition of UU if {Ui}i\{U_{i}\}_{i} is an at most countable family of Borel subsets of UU such that

μ(UiUi)=0,μ(UiUj)=0i,j;\mu\Big{(}U\setminus\bigcup_{i}U_{i}\Big{)}=0\,,\qquad\mu(U_{i}\cap U_{j})=0\quad\forall i,j\,; (1.45)

and we say that {Ui}i\{U_{i}\}_{i} is a monotone μ\mu-partition if, in addition to (1.45), we also have μ(Ui)μ(Ui+1)\mu(U_{i})\geq\mu(U_{i+1}) for every ii. When μ=n+1\mu=\mathcal{L}^{n+1} we replace “μ\mu-partition” with “Lebesgue partition”. When UU is a set of finite perimeter in n+1\mathbb{R}^{n+1}, we say that {Ui}i\{U_{i}\}_{i} is a Caccioppoli partition of UU if {Ui}i\{U_{i}\}_{i} is a Lebesgue partition of UU and each UiU_{i} is a set of finite perimeter in n+1\mathbb{R}^{n+1}: in this case we have

UniUi,2n(U(1)iUi)=in(U(1)Ui),\displaystyle\partial^{*}U\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\bigcup_{i}\partial^{*}U_{i}\,,\qquad 2\,\mathcal{H}^{n}\Big{(}U^{\scriptscriptstyle{(1)}}\cap\bigcup_{i}\partial^{*}U_{i}\Big{)}=\sum_{i}\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}U_{i})\,, (1.46)

see, e.g., [AFP00, Section 4.4]; moreover,

1#{i:xUi}2,xiUi,1\leq\#\Big{\{}i:x\in\partial^{*}U_{i}\Big{\}}\leq 2\,,\qquad\forall x\in\bigcup_{i}\partial^{*}U_{i}\,, (1.47)

thanks to (1.36) and to the fact that there cannot be three disjoint half-spaces in n+1\mathbb{R}^{n+1}.

2. Induced essential partitions (Theorem 1.2)

Given a Borel set SS, we say that a Lebesgue partition {Ui}i\{U_{i}\}_{i} of UU is induced by SS if, for each ii,

U(1)eUi is n-contained in S.\mbox{$U^{\scriptscriptstyle{(1)}}\cap\partial^{e}U_{i}$ is $\mathcal{H}^{n}$-contained in $S$}\,. (2.1)

We say that {Ui}i\{U_{i}\}_{i} is an essential partition of UU induced by SS if it is a Lebesgue partition of UU induced by SS such that, for each ii,

S does not essentially disconnect Ui.\mbox{$S$ does not essentially disconnect $U_{i}$}\,. (2.2)

The next theorem, which expands the statement of Theorem 1.2, shows that when n\mathcal{H}^{n}-finite sets uniquely determine induced essential partitions on sets of finite perimeter.

Theorem 2.1 (Induced essential partitions).

If Un+1U\subset\mathbb{R}^{n+1} is a bounded set of finite perimeter and Sn+1S\subset\mathbb{R}^{n+1} is a Borel set with n(SU(1))<\mathcal{H}^{n}(S\cap U^{\scriptscriptstyle{(1)}})<\infty, then there exists an essential partition {Ui}i\{U_{i}\}_{i} of UU induced by SS such that each UiU_{i} is a set of finite perimeter and

iP(Ui;U(1))2n(SU(1)).\sum_{i}P(U_{i};U^{\scriptscriptstyle{(1)}})\leq 2\,\mathcal{H}^{n}(S\cap U^{\scriptscriptstyle{(1)}})\,. (2.3)

Moreover: (a): if SS^{*} is a Borel set with n(SU(1))<\mathcal{H}^{n}(S^{*}\cap U^{\scriptscriptstyle{(1)}})<\infty, SS^{*} is n\mathcal{H}^{n}-contained in SS, {Vj}j\{V_{j}\}_{j} is a Lebesgue partition888Notice that here we are not requiring that SS^{*} does not essentially disconnect each VjV_{j}, i.e., we are not requiring that {Vj}j\{V_{j}\}_{j} is an essential partition induced by SS^{*}. This detail will be useful in the applications of this theorem. of UU induced by SS^{*}, and {Ui}i\{U_{i}\}_{i} is the essential partition of UU induced by SS, then

jVj is n-contained in iUi;\mbox{$\bigcup_{j}\,\partial^{*}V_{j}$ is $\mathcal{H}^{n}$-contained in $\bigcup_{i}\,\partial^{*}U_{i}$}\,; (2.4)

(b): if SS and SS^{*} are n\mathcal{H}^{n}-finite sets in U(1)U^{\scriptscriptstyle{(1)}}, and either999Here (S)\mathcal{R}(S) denotes the n\mathcal{H}^{n}-rectifiable part of SS. S=(S)S^{*}=\mathcal{R}(S) or SS^{*} is n\mathcal{H}^{n}-equivalent to SS, then SS and SS^{*} induce n+1\mathcal{L}^{n+1}-equivalent essential partitions of UU.

Proof of Theorem 1.2.

Immediate consequence of Theorem 2.1. ∎

The proof of Theorem 2.1 follows the main lines of the proof of [ACMM01, Theorem 1], which is indeed the case S=S=\varnothing of Theorem 2.1. We premise to this proof two lemmas that will find repeated applications in later sections too. To introduce the first lemma, we notice that, while it is evident that if SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} and SS is n\mathcal{H}^{n}-contained into some Borel set SS^{*}, then SS^{*} is also 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, however, it is not immediately clear if the rectifiable part (S)\mathcal{R}(S) of SS (which may not be n\mathcal{H}^{n}-equivalent to SS) retains the 𝒞\mathcal{C}-spanning property.

Lemma 2.2.

If 𝐖\mathbf{W} is compact, 𝒞\mathcal{C} is a spanning class for 𝐖\mathbf{W}, SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, and n  S\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S is a Radon measure in Ω\Omega, then (S)\mathcal{R}(S) is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}. Moreover, the sets T1T_{1} and T2T_{2} appearing in (1.12) are sets of finite perimeter.

Proof.

We make the following claim: if TT is open, T(1)nTT^{{\scriptscriptstyle{(1)}}}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}T, n  Z\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}Z is a Radon measure in an open neighborhood of TT, and ZZ essentially disconnects TT into {T1,T2}\{T_{1},T_{2}\}, then

T1 and T2 are of locally finite perimeter in T,\displaystyle\mbox{$T_{1}$ and $T_{2}$ are of locally finite perimeter in $T$}\,, (2.5)
(Z) essentially disconnects T into {T1,T2}.\displaystyle\mbox{$\mathcal{R}(Z)$ essentially disconnects $T$ into $\{T_{1},T_{2}\}$}\,. (2.6)

Indeed: Since TT is open, we trivially have TT(1)T\subset T^{{\scriptscriptstyle{(1)}}}, and hence TT is n\mathcal{H}^{n}-equivalent to T(1)T^{{\scriptscriptstyle{(1)}}}. Taking also into account that ZZ essentially disconnects TT into {T1,T2}\{T_{1},T_{2}\}, we thus find

TeT1eT2=nT(1)eT1eT2nZT(1)nZT.\displaystyle T\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}T^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}Z\cap T^{{\scriptscriptstyle{(1)}}}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}Z\cap T\,.

By Federer’s criterion (1.38) and the n\mathcal{H}^{n}-finiteness of ZZ in an open neighborhood of TT we deduce (2.5). By Federer’s theorem (1.37), eTi\partial^{e}T_{i} is (n  T)(\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}T)-equivalent to Ti\partial^{*}T_{i}, which combined with the n\mathcal{H}^{n}-equivalence of T(1)T^{{\scriptscriptstyle{(1)}}} and TT gives

eT1eT2T(1)=nT1T2T.\displaystyle\partial^{e}T_{1}\cap\partial^{e}T_{2}\cap T^{{\scriptscriptstyle{(1)}}}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\partial^{*}T_{1}\cap\partial^{*}T_{2}\cap T\,.

Since T1T2T\partial^{*}T_{1}\cap\partial^{*}T_{2}\cap T is n\mathcal{H}^{n}-rectifiable and eT1eT2T(1)nZ\partial^{e}T_{1}\cap\partial^{e}T_{2}\cap T^{{\scriptscriptstyle{(1)}}}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}Z, we conclude that n(eT1eT2T(1)𝒫(Z))=0\mathcal{H}^{n}(\partial^{e}T_{1}\cap\partial^{e}T_{2}\cap T^{{\scriptscriptstyle{(1)}}}\cap\mathcal{P}(Z))=0. Hence,

eT1eT2T(1)n(Z),\partial^{e}T_{1}\cap\partial^{e}T_{2}\cap T^{{\scriptscriptstyle{(1)}}}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\mathcal{R}(Z)\,,

and (2.6) follows.

To prove the lemma: Let (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}), let JJ be of full measure such that (A.1) holds for every sJs\in J, so that, for every sJs\in J one finds that for n\mathcal{H}^{n}-a.e. xT[s]x\in T[s] there is a partition {T1,T2}\{T_{1},T_{2}\} of TT with xeT1eT2x\in\partial^{e}T_{1}\cap\partial^{e}T_{2} and such that ST[s]S\cup T[s] essentially disconnects TT into {T1,T2}\{T_{1},T_{2}\}. By applying the claim with Z=ST[s]Z=S\cup T[s], we see that (ST[s])\mathcal{R}(S\cup T[s]) essentially disconnects TT into {T1,T2}\{T_{1},T_{2}\}, and that T1T_{1} and T2T_{2} have locally finite perimeter in TT. On noticing that (ST[s])\mathcal{R}(S\cup T[s]) is n\mathcal{H}^{n}-equivalent to (S)T[s]\mathcal{R}(S)\cup T[s], we conclude the proof. ∎

The second lemma is just a simple compactness statement for finite perimeter partitions.

Lemma 2.3 (Compactness for partitions by sets of finite perimeter).

If UU is a bounded open set and {{Uij}i=1}j=1\{\{U_{i}^{j}\}_{i=1}^{\infty}\}_{j=1}^{\infty} is a sequence of Lebesgue partitions of UU into sets of finite perimeter such that

supji=1P(Uij)<,\displaystyle\sup_{j}\,\sum_{i=1}^{\infty}P(U_{i}^{j})<\infty\,, (2.7)

then, up to extracting a subsequence, there exists a Lebesgue partition {Ui}i\{U_{i}\}_{i\in\mathbb{N}} of UU such that for every ii and every AUA\subset U open,

limj|UijΔUi|=0,P(Ui;A)lim infjP(Uij;A).\displaystyle\lim_{j\to\infty}|U_{i}^{j}\Delta U_{i}|=0\,,\qquad P(U_{i};A)\leq\liminf_{j\to\infty}P(U_{i}^{j};A)\,. (2.8)

Moreover,

limilim supjk=i+1|Ukj|s=0,s(nn+1,1).\displaystyle\lim_{i\to\infty}\limsup_{j\to\infty}\sum_{k=i+1}^{\infty}|U^{j}_{k}|^{s}=0\,,\qquad\forall s\in\Big{(}\frac{n}{n+1},1\Big{)}\,. (2.9)
Proof.

Up to a relabeling we can assume each {Uij}i\{U_{i}^{j}\}_{i} is monotone. By (2.7) and the boundedness of UU, a diagonal argument combined with standard lower semicontinuity and compactness properties of sets of finite perimeter implies that we can find a not relabeled subsequence in jj and a family {Ui}i\{U_{i}\}_{i} of Borel subsets of UU with |Ui||Ui+1||U_{i}|\geq|U_{i+1}| and |UiUj|=0|U_{i}\cap U_{j}|=0 for every iji\neq j, such that (2.8) holds. We are thus left to prove (2.9) and

|Ui=1Ui|=0.\displaystyle\Big{|}U\setminus\bigcup_{i=1}^{\infty}U_{i}\Big{|}=0\,. (2.10)

We start by noticing that for each ii there is J(i)J(i)\in\mathbb{N} such that |Ukj|2|Uk||U_{k}^{j}|\leq 2\,|U_{k}| for every jJ(i)j\geq J(i) and 1ki1\leq k\leq i. Therefore if ki+1k\geq i+1 and jJ(i)j\geq J(i) we find |Ukj||Uij|2|Ui||U_{k}^{j}|\leq|U_{i}^{j}|\leq 2\,|U_{i}|, so that, if jJ(i)j\geq J(i),

k=i+1|Ukj|sC(n)k=i+1P(Ukj)|Ukj|s(n/(n+1))C|Ui|s(n/(n+1)),\displaystyle\sum_{k=i+1}^{\infty}|U_{k}^{j}|^{s}\leq C(n)\,\sum_{k=i+1}^{\infty}P(U_{k}^{j})|U_{k}^{j}|^{s-(n/(n+1))}\leq C\,|U_{i}|^{s-(n/(n+1))}\,, (2.11)

where we have also used the isoperimetric inequality and (2.7). Since |Ui|0|U_{i}|\to 0 as ii\to\infty (indeed, i|Ui||U|<\sum_{i}|U_{i}|\leq|U|<\infty), (2.11) implies (2.9). To prove (2.10), we notice that if we set M=|UiUi|M=|U\setminus\cup_{i}U_{i}|, and we assume that MM is positive, then up to further increasing the value of J(i)J(i) we can require that

|Ukj||Uk|+M2k+2,1ki,jJ(i),\displaystyle|U_{k}^{j}|\leq|U_{k}|+\frac{M}{2^{k+2}}\,,\qquad\forall 1\leq k\leq i\,,\,\forall j\geq J(i)\,, (2.12)

(in addition to |Ukj|2|Uk||U_{k}^{j}|\leq 2\,|U_{k}|). By (2.12) we obtain that, if jJ(i)j\geq J(i), then

|U|k=i+1|Ukj|=k=1i|Ukj|k=1i|Uk|+M2k+2|U|M+k=1iM2k+2|U|M4.\displaystyle|U|-\sum_{k=i+1}^{\infty}|U_{k}^{j}|=\sum_{k=1}^{i}|U_{k}^{j}|\leq\sum_{k=1}^{i}|U_{k}|+\frac{M}{2^{k+2}}\leq|U|-M+\sum_{k=1}^{i}\frac{M}{2^{k+2}}\leq|U|-\frac{M}{4}\,. (2.13)

Rearranging (2.13) and using the sub-additivity of zzsz\mapsto z^{s} we conclude that

(M/4)sk=i+1|Ukj|s.\displaystyle(M/4)^{s}\leq\sum_{k=i+1}^{\infty}|U_{k}^{j}|^{s}\,.

We obtain a contradiction with M>0M>0 by letting ii\to\infty and by using (2.9). ∎

Proof of Theorem 2.1.

Let 𝒰(S)\mathcal{U}(S) be the set of all the monotone Lebesgue partitions of UU induced by SS. We notice that 𝒰(S)\mathcal{U}(S)\neq\varnothing, since 𝒰(S)\mathcal{U}(S) contains the trivial partition with U1=UU_{1}=U and Ui=U_{i}=\varnothing if i2i\geq 2. If Ui{Ui}iU_{i}\in\{U_{i}\}_{i} for {Ui}i𝒰(S)\{U_{i}\}_{i}\in\mathcal{U}(S), then eUi\partial^{e}U_{i} is n\mathcal{H}^{n}-contained in eU(U(1)S)\partial^{e}U\cup(U^{(1)}\cap S), which, by Federer’s criterion applied to UU and n(SU(1))<\mathcal{H}^{n}(S\cap U^{\scriptscriptstyle{(1)}})<\infty, has finite n\mathcal{H}^{n}-measure; it follows then that UiU_{i} is a set of finite perimeter due to Federer’s criterion. We now fix s(n/(n+1),1)s\in(n/(n+1),1), and consider a maximizing sequence {{Uij}i}j\{\{U_{i}^{j}\}_{i}\}_{j} for

m=max{i=1|Ui|s:{Ui}i𝒰(S)}.m=\max\Big{\{}\sum_{i=1}^{\infty}|U_{i}|^{s}:\{U_{i}\}_{i}\in\mathcal{U}(S)\Big{\}}\,.

By standard arguments concerning reduced boundaries of disjoint sets of finite perimeter (see, e.g. [Mag12, Chapter 16]), we deduce from (2.1) that for every jj,

i=1n Uij\displaystyle\sum_{i=1}^{\infty}\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}\partial^{*}U_{i}^{j} =\displaystyle= i=1n (UijU(1))+i=1n (UijU)\displaystyle\sum_{i=1}^{\infty}\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}U_{i}^{j}\cap U^{{\scriptscriptstyle{(1)}}})+\sum_{i=1}^{\infty}\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}U_{i}^{j}\cap\partial^{*}U) (2.14)
\displaystyle\leq 2n (SU(1))+n U.\displaystyle 2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(S\cap U^{\scriptscriptstyle{(1)}})+\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}\partial^{*}U\,.

Also, due to the sub-additivity of zzsz\mapsto z^{s} and the general fact that e(AB)eAeB\partial^{e}(A\cap B)\subset\partial^{e}A\cup\partial^{e}B, we can refine {Uij}i\{U^{j}_{i}\}_{i} by replacing each UijU_{i}^{j} with the disjoint family

{UijUk:k1,1<j},\big{\{}U_{i}^{j}\cap U^{\ell}_{k}:k\geq 1\,,1\leq\ell<j\big{\}}\,,

thus obtaining a new sequence in 𝒰(S)\mathcal{U}(S) which is still maximizing for mm. As a consequence of this remark, we can assume without loss of generality that the considered maximizing sequence {{Uij}i}j\{\{U_{i}^{j}\}_{i}\}_{j} for mm has the additional property that

UiUijUiUij+1,j.\displaystyle U\cap\bigcup_{i}\partial^{*}U_{i}^{j}\subset U\cap\bigcup_{i}\partial^{*}U_{i}^{j+1}\,,\qquad\forall j\,. (2.15)

Thanks to (2.14) we can apply Lemma 2.3 and, up to extracting a subsequence in jj, we can find a Lebesgue partition {Ui}i\{U_{i}\}_{i\in\mathbb{N}} of UU by sets of finite perimeter which satisfies (2.8) and (2.9). Moreover, after taking a subsequence, we may assume that n  Uijμi\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}\partial^{*}U_{i}^{j}\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}\mu_{i} for some Radon measures μi\mu_{i} such that n  Uiμi\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}\partial^{*}U_{i}\leq\mu_{i} [Mag12, Prop. 12.15]. Therefore, by (2.8), Federer’s theorem for reduced boundaries, and by (2.1) for {Uij}i\{U^{j}_{i}\}_{i}, we see that

n\displaystyle\mathcal{H}^{n}  (U)+i=1n (UiU(1))=i=1n (Ui)wlimji=1n (Uij)\displaystyle\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}U)+\sum_{i=1}^{\infty}\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}U_{i}\cap U^{\scriptscriptstyle{(1)}})=\sum_{i=1}^{\infty}\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}U_{i})\leq{\rm w}^{*}\lim_{j\to\infty}\sum_{i=1}^{\infty}\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}U_{i}^{j})
=wlimjn (U)+i=1n (eUijU(1))n (U)+2n (SU(1)).\displaystyle={\rm w}^{*}\lim_{j\to\infty}\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}U)+\sum_{i=1}^{\infty}\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{e}U_{i}^{j}\cap U^{\scriptscriptstyle{(1)}})\leq\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}U)+2\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(S\cap U^{\scriptscriptstyle{(1)}})\,.

By subtracting n  (U)\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}U) from both sides, we deduce (2.3).

We now show, first, that {Ui}i𝒰(S)\{U_{i}\}_{i}\in\mathcal{U}(S) (i.e., we check the validity of (2.1) on {Ui}i\{U_{i}\}_{i}), and then that SS does not essentially disconnect any of the UiU_{i}. This will complete the proof of the first part of the statement.

To prove that U(1)eUinSU^{\scriptscriptstyle{(1)}}\cap\partial^{e}U_{i}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}S, let us introduce the n\mathcal{H}^{n}-rectifiable set S0S_{0} defined by

S0=U(1)i,jUij.\displaystyle S_{0}=U^{\scriptscriptstyle{(1)}}\cap\bigcup_{i,j}\partial^{*}U_{i}^{j}\,. (2.16)

By {Uij}i𝒰(S)\{U_{i}^{j}\}_{i}\in\mathcal{U}(S), S0S_{0} is contained into SS modulo n\mathcal{H}^{n}-null sets. Therefore, in order to prove (2.1) it will be enough to show that

U(1)UinS0,i.\displaystyle U^{{\scriptscriptstyle{(1)}}}\cap\partial^{*}U_{i}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}S_{0}\,,\qquad\forall i\,. (2.17)

Should this not be the case, it would be n(U(1)UiS0)>0\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}U_{i}\setminus S_{0})>0 for some ii. We could thus pick xU(1)Uix\in U^{\scriptscriptstyle{(1)}}\cap\partial^{*}U_{i} such that

θn(n (U(1)UiS0))(x)=1.\displaystyle\theta^{n}(\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}U_{i}\setminus S_{0}))(x)=1\,. (2.18)

Since θn(n  Ui)(x)=1\theta^{n}(\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}\partial^{*}U_{i})(x)=1 and S0U(1)S_{0}\subset U^{\scriptscriptstyle{(1)}} this implies n(S0Br(x))=o(rn)\mathcal{H}^{n}(S_{0}\cap B_{r}(x))={\rm o}(r^{n}), while UiUi(1/2)\partial^{*}U_{i}\subset U_{i}^{\scriptscriptstyle{(1/2)}} gives |UiBr(x)|=(ωn+1/2)rn+1+o(rn+1)|U_{i}\cap B_{r}(x)|=(\omega_{n+1}/2)\,r^{n+1}+{\rm o}(r^{n+1}). Therefore, given δ>0\delta>0 we can find r>0r>0 such that

n(S0Br(x))<δrn,min{|UiBr(x)|,|UiBr(x)|}(ωn+12δ)rn+1,\mathcal{H}^{n}(S_{0}\cap B_{r}(x))<\delta\,r^{n}\,,\qquad\min\big{\{}|U_{i}\cap B_{r}(x)|,|U_{i}\setminus B_{r}(x)|\big{\}}\geq\Big{(}\frac{\omega_{n+1}}{2}-\delta\Big{)}\,r^{n+1}\,,

and then exploit the relative isoperimetric inequality and (2.8) to conclude that

c(n)[(ωn+12δ)rn+1]n/(n+1)\displaystyle c(n)\,\Big{[}\Big{(}\frac{\omega_{n+1}}{2}-\delta\Big{)}\,r^{n+1}\Big{]}^{n/(n+1)} \displaystyle\leq P(Ui;Br(x))lim infjP(Uij;Br(x))\displaystyle P(U_{i};B_{r}(x))\leq\liminf_{j\to\infty}P(U_{i}^{j};B_{r}(x))
\displaystyle\leq n(S0Br(x))δrn,\displaystyle\mathcal{H}^{n}(S_{0}\cap B_{r}(x))\leq\delta\,r^{n}\,,

where in the next to last inequality we have used the definition (2.16) of S0S_{0}. Choosing δ>0\delta>0 small enough we reach a contradiction, thus deducing that {Ui}i𝒰(S)\{U_{i}\}_{i}\in\mathcal{U}(S).

Taking into account the subadditivity of zzsz\mapsto z^{s}, in order to prove that SS does not essentially disconnect any UiU_{i} it is sufficient to show that {Ui}i\{U_{i}\}_{i} is a maximizer of mm. To see this, we notice that |UijΔUi|0|U_{i}^{j}\Delta U_{i}|\to 0 as jj\to\infty implies

m=limji=1k|Uij|s+i=k+1|Uij|s=i=1k|Ui|s+limji=k+1|Uij|s,m=\lim_{j\to\infty}\sum_{i=1}^{k}|U_{i}^{j}|^{s}+\sum_{i=k+1}^{\infty}|U_{i}^{j}|^{s}=\sum_{i=1}^{k}|U_{i}|^{s}+\lim_{j\to\infty}\sum_{i=k+1}^{\infty}|U_{i}^{j}|^{s}\,,

so that, letting kk\to\infty and exploiting (2.9), we conclude that

m=i=1|Ui|s.m=\sum_{i=1}^{\infty}|U_{i}|^{s}\,. (2.19)

This completes the proof of the first part of the statement (existence of essential partitions).

Let now SS, SS^{*}, {Ui}i\{U_{i}\}_{i}, and {Uj}j\{U_{j}^{*}\}_{j} be as in statement (a) – that is, SS^{*} is n\mathcal{H}^{n}-contained in SS, {Ui}i\{U_{i}\}_{i} is an essential partition of UU induced by SS, and, for every jj, {Uj}j\{U_{j}^{*}\}_{j} is a Lebesgue partition of UU induced by SS^{*} – and set Z=iUiZ=\cup_{i}\partial^{*}U_{i} and Z=jUjZ^{*}=\cup_{j}\partial^{*}U_{j}^{*}. Arguing by contradiction with (2.4), let us assume n(ZZ)>0\mathcal{H}^{n}(Z^{*}\setminus Z)>0. By the definition of Lebesgue partition we have that ZU(1)Z\setminus U^{\scriptscriptstyle{(1)}} and ZU(1)Z^{*}\setminus U^{\scriptscriptstyle{(1)}} are both n\mathcal{H}^{n}-equivalent to U\partial^{*}U. Therefore we have n((ZZ)U(1))>0\mathcal{H}^{n}((Z^{*}\setminus Z)\cap U^{\scriptscriptstyle{(1)}})>0. Since U(1)U^{\scriptscriptstyle{(1)}} is n\mathcal{H}^{n}-equivalent to the union of the {Ui(1)Ui}iI\{U_{i}^{{\scriptscriptstyle{(1)}}}\cup\partial^{*}U_{i}\}_{i\in I} we can find iIi\in I and jJj\in J such that n(Ui(1)Uj)>0\mathcal{H}^{n}(U_{i}^{{\scriptscriptstyle{(1)}}}\cap\partial^{*}U_{j}^{*})>0. This implies that both (UiUj)(1/2)(U_{i}\cap U_{j}^{*})^{\scriptscriptstyle{(1/2)}} and (UiUj)(1/2)(U_{i}\setminus U_{j}^{*})^{\scriptscriptstyle{(1/2)}} are non-empty, and thus that {UjUi,UiUj}\{U_{j}^{*}\cap U_{i},U_{i}\setminus U_{j}^{*}\} is a non-trivial Borel partition of UiU_{i}. Since

Ui(1)e(UjUi)nU(1)UjnS,U_{i}^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}(U_{j}^{*}\cap U_{i})\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}U^{\scriptscriptstyle{(1)}}\cap\partial^{*}U_{j}^{*}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}S^{*}\,,

we conclude that SS^{*} is essentially disconnecting UiU_{i}, against the fact that SS is not essentially disconnecting UiU_{i} and the fact that SS^{*} is n\mathcal{H}^{n}-contained in SS.

We finally prove statement (b). Let {Ui}iI\{U_{i}\}_{i\in I}, and {Uj}jJ\{U_{j}^{*}\}_{j\in J} be essential partitions of UU induced by SS and SS^{*} respectively. Given iIi\in I such that |Ui|>0|U_{i}|>0, there is at least one jJj\in J such that |UiUj|>0|U_{i}\cap U_{j}^{*}|>0. We claim that it must be |UiUj|=0|U_{i}\setminus U_{j}^{*}|=0. Should this not be the case, Uj\partial^{*}U_{j}^{*} would be essentially disconnecting UiU_{i}, thus implying that SS^{*} (which contains Uj\partial^{*}U_{j}^{*}) is essentially disconnecting UiU_{i}. Now, either because we are assuming that SS^{*} is n\mathcal{H}^{n}-equivalent to SS, or because we are assuming that S=(S)S^{*}=\mathcal{R}(S) and we have Lemma 2.2, the fact that SS^{*} is essentially disconnecting UiU_{i} implies that SS is essentially disconnecting UiU_{i}, a contradiction. Having proved the claim, for each iIi\in I with |Ui|>0|U_{i}|>0 there is a unique σ(i)J\sigma(i)\in J such that |UiΔUσ(j)|=0|U_{i}\Delta U_{\sigma(j)}^{*}|=0. This completes the proof. ∎

3. Homotopic spanning on generalized soap films (Theorem 1.3)

The goal of this section is proving Theorem 1.3, and, actually, to obtain an even more general result. Let us recall that the objective of Theorem 1.3 was to reformulate the homotopic spanning property for a Borel set SS, in the case when SS is locally n\mathcal{H}^{n}-finite, in terms of unions of boundaries of induced essential partitions. We shall actually need this kind of characterization also for sets SS of the more general form S=KE(1)S=K\cup E^{\scriptscriptstyle{(1)}}, where (K,E)𝒦B(K,E)\in\mathcal{K}_{\rm B}. For an illustration of the proposed characterization of homotopic spanning on this type of sets, see

Refer to caption
𝐖\mathbf{W}TTKKT[s]T[s]EE(a)(a)(b)(b)U1U_{1}U2U_{2}U3U_{3}U4U_{4}U5U_{5}U3U_{3}
Figure 3.1. In panel (a) we have depicted a pair (K,E)(K,E) where EE is a tube inside TT and KK consists of the union of the boundary of EE and the non-spanning set SS of Figure 1.6-(a). Notice that KK is not 𝒞\mathcal{C}-spanning, if we see things from the point of view of Definition A, since it misses every loop γ\gamma contained in the interior of EE; while, of course, KEK\cup E is 𝒞\mathcal{C}-spanning because EE has been added. In panel (b) we have depicted the essential partition {Ui}i=15\{U_{i}\}_{i=1}^{5} of TT induced by KT[s]K\cup T[s]. Notice that E=U1E=U_{1}, therefore no UiUj\partial^{*}U_{i}\cap\partial^{*}U_{j} 1\mathcal{H}^{1}-containis T[s]ET[s]\cap E. In particular, T[s]ET[s]\cap E (which 1\mathcal{H}^{1}-equivalent to T[s]E(0)T[s]\setminus E^{\scriptscriptstyle{(0)}}) is not 1\mathcal{H}^{1}-contained in UBEP(KT[s];T){\rm UBEP}(K\cup T[s];T), and we see again, this time from the point of view of Definition B as reformulated in Theorem 1.3, that KK is not 𝒞\mathcal{C}-spanning. As stated in Theorem 3.1, from the viewpoint of Definition B it is only the 1\mathcal{H}^{1}-containment of T[s]E(0)T[s]\cap E^{\scriptscriptstyle{(0)}} into UBEP(KT[s];T){\rm UBEP}(K\cup T[s];T) that establishes the 𝒞\mathcal{C}-spanning property of KEK\cup E: and this 1\mathcal{H}^{1}-containment indeed holds, since T[s]E(0)=T[s]cl(E)T[s]\cap E^{\scriptscriptstyle{(0)}}=T[s]\setminus\mathrm{cl}\,(E) is 1\mathcal{H}^{1}-contained in the union of U2U3\partial^{*}U_{2}\cap\partial^{*}U_{3} and U4U5\partial^{*}U_{4}\cap\partial^{*}U_{5}.

Figure 3.1.

Theorem 3.1 (Homotopic spanning for generalized soap films).

If 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} is a closed set in n+1\mathbb{R}^{n+1}, 𝒞\mathcal{C} is a spanning class for 𝐖\mathbf{W}, KK is a Borel set locally n\mathcal{H}^{n}-finite in Ω\Omega, and EE is of locally finite perimeter in Ω\Omega such that ΩE\Omega\cap\partial^{*}E is n\mathcal{H}^{n}-contained in KK, then the set

S=(K)E(1)\displaystyle S=\mathcal{R}(K)\cup E^{{\scriptscriptstyle{(1)}}} (3.1)

is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} if and only if, for every (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}) and 1\mathcal{H}^{1}-a.e. sSS1s\in\SS^{1},

T[s]E(0) is n-contained in UBEP(KT[s];T).\displaystyle\mbox{$T[s]\cap E^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in ${\rm UBEP}(K\cup T[s];T)$}\,. (3.2)
Remark 3.2.

An immediate corollary of Theorem 3.1 is that if KK is n\mathcal{H}^{n}-finite and (K,E)𝒦B(K,E)\in\mathcal{K}_{\rm B} then KE(1)K\cup E^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} if and only if (K)E(1)\mathcal{R}(K)\cup E^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}. Indeed, (KT[s])=(K)T[s]\mathcal{R}(K\cup T[s])=\mathcal{R}(K)\cup T[s], so that, by (1.13), UBEP(KT[s])=UBEP((K)T[s]){\rm UBEP}(K\cup T[s])={\rm UBEP}(\mathcal{R}(K)\cup T[s]).

Proof of Theorem 1.3.

This is Theorem 3.1 with E=E=\varnothing. ∎

Proof of Theorem 3.1.

Step one: We prove the following claim: If SS essentially disconnects GG into {G1,G2}\{G_{1},G_{2}\} and HGH\subset G satisfies

min{|HG1|,|HG2|}>0,\displaystyle\min\{|H\cap G_{1}|\,,\,|H\cap G_{2}|\}>0\,, (3.3)

then SS essentially disconnects HH into HG1H\cap G_{1} and HG2H\cap G_{2}. Indeed, if xH(1)x\in H^{{\scriptscriptstyle{(1)}}}, then xe(HGi)x\in\partial^{e}(H\cap G_{i}) if and only if xeGix\in\partial^{e}G_{i} (i=1,2i=1,2). Hence H(1)e(G1H)H(1)eG1G(1)eG1H^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}(G_{1}\cap H)\subset H^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}G_{1}\subset G^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}G_{1}, which, by (3.3) and our assumption on SS and GG, gives the desired conclusion.

Step two: Taking from now on SS, KK and EE as in the statement we preliminary notice that if (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}), sSS1s\in\SS^{1}, and {Ui}i\{U_{i}\}_{i} is the essential partition of TT induced by ((K)T[s])(\mathcal{R}(K)\cup T[s]), then

TEnTiUi.\displaystyle T\cap\partial^{*}E\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}T\cap\bigcup_{i}\partial^{*}U_{i}\,. (3.4)

Indeed, since ΩE\Omega\cap\partial^{*}E is n\mathcal{H}^{n}-contained in (K)\mathcal{R}(K), if a Borel set GG is such that |GE||GE|>0|G\cap E|\,|G\setminus E|>0 then, by step one, (K)\mathcal{R}(K) essentially disconnects GG. In particular, since, for each ii, (K)T[s]\mathcal{R}(K)\cup T[s] does not essentially disconnect UiU_{i}, we find that, for each ii,

either Ui(1)E(0)or Ui(1)E(1).\mbox{either $U_{i}^{{\scriptscriptstyle{(1)}}}\subset E^{\scriptscriptstyle{(0)}}$}\qquad\mbox{or $U_{i}^{{\scriptscriptstyle{(1)}}}\subset E^{{\scriptscriptstyle{(1)}}}$}\,. (3.5)

Clearly, (3.5) immediately implies (3.4).

Step three: We prove the “only if” part of the statement, that is, given (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}) and s𝕊1s\in\mathbb{S}^{1}, we assume that

for n-a.e. xT[s],\displaystyle\mbox{for $\mathcal{H}^{n}$-a.e. $x\in T[s]$}\,, (3.6)
 a partition {T1,T2} of T with xeT1eT2,\displaystyle\mbox{$\exists$ a partition $\{T_{1},T_{2}\}$ of $T$ with $x\in\partial^{e}T_{1}\cap\partial^{e}T_{2}$}\,,
and s.t. (K)E(1)T[s] essentially disconnects T into {T1,T2},\displaystyle\mbox{and s.t. $\mathcal{R}(K)\cup E^{{\scriptscriptstyle{(1)}}}\cup T[s]$ essentially disconnects $T$ into $\{T_{1},T_{2}\}$}\,,

and then prove that

T[s]E(0) is n-contained in iUi,\displaystyle\mbox{$T[s]\cap E^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in $\bigcup_{i}\partial^{*}U_{i}$}\,, (3.7)

where {Ui}i\{U_{i}\}_{i} is the essential partition of TT induced by (K)T[s]\mathcal{R}(K)\cup T[s]. To this end, arguing by contradiction, we suppose that for some s𝕊1s\in\mathbb{S}^{1}, there is GT[s]E(0)G\subset T[s]\cap E^{\scriptscriptstyle{(0)}} with n(G)>0\mathcal{H}^{n}(G)>0 and such that GiUi=G\cap\cup_{i}\partial^{*}U_{i}=\varnothing. In particular, there is an index ii such that n(GUi(1))>0\mathcal{H}^{n}(G\cap U_{i}^{{\scriptscriptstyle{(1)}}})>0, which, combined with (3.5) and GE(0)G\subset E^{\scriptscriptstyle{(0)}}, implies

Ui(1)E(0).U_{i}^{{\scriptscriptstyle{(1)}}}\subset E^{\scriptscriptstyle{(0)}}\,. (3.8)

Now by (3.6) and n(GUi(1))>0\mathcal{H}^{n}(G\cap U^{{\scriptscriptstyle{(1)}}}_{i})>0, we can choose xGUi(1)x\in G\cap U_{i}^{{\scriptscriptstyle{(1)}}} such that (K)E(1)T[s]\mathcal{R}(K)\cup E^{{\scriptscriptstyle{(1)}}}\cup T[s] essentially disconnects TT into some {T1,T2}\{T_{1},T_{2}\} such that xeT1eT2x\in\partial^{e}T_{1}\cap\partial^{e}T_{2}. Then, {UiT1,UiT2}\{U_{i}\cap T_{1},U_{i}\cap T_{2}\} is a non-trivial partition of UiU_{i}, so that, by step one and (3.8), (K)T[s]\mathcal{R}(K)\cup T[s] essentially disconnects UiU_{i} into {UiT1,UiT2}\{U_{i}\cap T_{1},U_{i}\cap T_{2}\}. This contradicts the defining property (2.2) of essential partitions, and concludes the proof.

Step four: We prove the “if” part of the statement. More precisely, given (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}) and s𝕊1s\in\mathbb{S}^{1}, we assume that (3.7) holds at ss, and then proceed to prove that (3.6) holds at ss. We first notice that, since {E(1),E(0),E}\{E^{{\scriptscriptstyle{(1)}}},E^{\scriptscriptstyle{(0)}},\partial^{*}E\} is a partition of Ω\Omega modulo n\mathcal{H}^{n}, it is enough to prove (3.6) for n\mathcal{H}^{n}-a.e. xT[s](E(1)E(0)E)x\in T[s]\cap(E^{{\scriptscriptstyle{(1)}}}\cup E^{\scriptscriptstyle{(0)}}\cup\partial^{*}E).

If xT[s]Ex\in T[s]\cap\partial^{*}E, then by letting T1=TET_{1}=T\cap E and T2=TET_{2}=T\setminus E we obtain a partition of TT such that xTE=TT1T2eT1eT2x\in T\cap\partial^{*}E=T\cap\partial^{*}T_{1}\cap\partial^{*}T_{2}\subset\partial^{e}T_{1}\cap\partial^{e}T_{2}, and such that E\partial^{*}E essentially disconnects TT into {T1,T2}\{T_{1},T_{2}\}. Since ΩE\Omega\cap\partial^{*}E is n\mathcal{H}^{n}-contained in (K)\mathcal{R}(K), we deduce (3.6).

If xT[s]E(0)x\in T[s]\cap E^{\scriptscriptstyle{(0)}}, then, thanks to (3.7) and denoting by {Ui}i\{U_{i}\}_{i} the essential partition of TT induced by ((K)T[s])(\mathcal{R}(K)\cup T[s]), there is an index ii such that xTUix\in T\cap\partial^{*}U_{i}. Setting T1=UiT_{1}=U_{i} and T2=TUiT_{2}=T\setminus U_{i}, we have that TUiT\cap\partial^{*}U_{i} (which contains xx) is in turn contained into eT1eT2T\partial^{e}T_{1}\cap\partial^{e}T_{2}\cap T. Since the latter set is non-empty, {T1,T2}\{T_{1},T_{2}\} is a non-trivial partition of TT. Moreover, by definition of essential partition,

T(1)eT1eT2=TeUin(K)T[s],\displaystyle T^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}=T\cap\partial^{e}U_{i}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\mathcal{R}(K)\cup T[s]\,,

so that (K)T[s]\mathcal{R}(K)\cup T[s] essentially disconnects TT, and (3.6) holds.

Finally, if xT[s]E(1)x\in T[s]\cap E^{{\scriptscriptstyle{(1)}}}, we let s1=ss_{1}=s, pick s2ss_{2}\neq s, denote by {I1,I2}\{I_{1},I_{2}\} the partition of SS1\SS^{1} defined by {s1,s2}\{s_{1},s_{2}\}, and set

T1=Φ(I1×B1n)E,T2=Φ(I2×B1n)(Φ(I1×B1n)E).T_{1}=\Phi(I_{1}\times B_{1}^{n})\cap E\,,\qquad T_{2}=\Phi(I_{2}\times B_{1}^{n})\cup\,\Big{(}\Phi(I_{1}\times B_{1}^{n})\setminus E\Big{)}\,.

This is a Borel partition of TT, and using the fact that xE(1)x\in E^{{\scriptscriptstyle{(1)}}}, we compute

|T1Br(x)|=|Φ(I1×B1n)EBr(x)|=|Φ(I1×B1n)Br(x)|+o(rn+1)=|Br(x)|2+o(rn+1).|T_{1}\cap B_{r}(x)|=|\Phi(I_{1}\times B_{1}^{n})\cap E\cap B_{r}(x)|=|\Phi(I_{1}\times B_{1}^{n})\cap B_{r}(x)|+{\rm o}(r^{n+1})=\frac{|B_{r}(x)|}{2}+{\rm o}(r^{n+1})\,.

Therfore xeT1eT2x\in\partial^{e}T_{1}\cap\partial^{e}T_{2}, and by standard facts about reduced boundaries [Mag12, Chapter 16],

eT1eT2T(1)nT1T(1)n(E((T[s1]T[s2])E(1)))T(1).\partial^{e}T_{1}\cap\partial^{e}T_{2}\cap T^{{\scriptscriptstyle{(1)}}}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\partial^{*}T_{1}\cap T^{{\scriptscriptstyle{(1)}}}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\big{(}\partial^{*}E\cup((T[s_{1}]\cup T[s_{2}])\cap E^{{\scriptscriptstyle{(1)}}})\big{)}\cap T^{{\scriptscriptstyle{(1)}}}\,.

Since ΩE\Omega\cap\partial^{*}E is n\mathcal{H}^{n}-contained in (K)\mathcal{R}(K), we have shown (3.6). ∎

4. The fundamental closure theorem for homotopic spanning conditions

In Theorem 1.3 and Theorem 3.1 we have presented two reformulations of the homotopic spanning condition in terms of n\mathcal{H}^{n}-containment into union of boundaries of essential partitions. The goal of this section is discussing the closure of such reformulations, and provide a statement (Theorem 4.1 below) which will lie at the heart of the closure theorems proved in Section 5.

Theorem 4.1 (Basic closure theorem for homotopic spanning).

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be closed and let 𝒞\mathcal{C} be a spanning class for 𝐖\mathbf{W}. Let us assume that:

(a): KjK_{j} are n\mathcal{H}^{n}-finite Borel subsets of Ω\Omega with n  Kjμ\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}K_{j}\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}\mu as Radon measures in Ω\Omega;

(b): (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}), {sj}j\{s_{j}\}_{j} is a sequence in SS1\SS^{1} with sjs0s_{j}\to s_{0} as jj\to\infty;

(c): if {Uij}i\{U_{i}^{j}\}_{i} denotes the essential partition of TT induced by KjT[sj]K_{j}\cup T[s_{j}], then there is a limit partition {Ui}i\{U_{i}\}_{i} of {Uij}i\{U_{i}^{j}\}_{i} in the sense of (2.8) in Lemma 2.3;

Under these assumptions, if μ(T[s0])=0\mu(T[s_{0}])=0, Fj,FΩF_{j},F\subset\Omega are sets of finite perimeter with FjFF_{j}\to F as jj\to\infty and such that, for every jj, ΩFj\Omega\cap\partial^{*}F_{j} is n\mathcal{H}^{n}-contained in KjK_{j} and

T[sj]Fj(0) is n-contained in Kj,\mbox{$T[s_{j}]\cap F_{j}^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in $K^{*}_{j}$}\,, (4.1)

then

T[s0]F(0) is n-contained in K,\mbox{$T[s_{0}]\cap F^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in $K^{*}$}\,, (4.2)

where we have set

Kj=UBEP(KjT[sj];T)=TiUij,K=TiUi.K_{j}^{*}={\rm UBEP}(K_{j}\cup T[s_{j}];T)=T\cap\bigcup_{i}\partial^{*}U_{i}^{j}\,,\qquad K^{*}=T\cap\bigcup_{i}\partial^{*}U_{i}\,. (4.3)
Remark 4.2.

Notice that {Ui}i\{U_{i}\}_{i} may fail to be the essential partition of TT induced by KK^{*} (which is the “optimal” choice of a Borel set potentially inducing {Ui}i\{U_{i}\}_{i} on TT): indeed, some of the sets UiU_{i} may fail to be essentially connected, even though UijUiU_{i}^{j}\to U_{i} as jj\to\infty and every UijU_{i}^{j}, as an element of an essential partition, is necessarily essentially connected; see

Refer to caption
TTFjF_{j}T[s0]T[s_{0}]V6V_{6}T[sj]T[s_{j}]KK^{*}U1jU_{1}^{j}U1jU_{1}^{j}U2jU_{2}^{j}U3jU_{3}^{j}U4jU_{4}^{j}U5jU_{5}^{j}V1V_{1}V1V_{1}V2V_{2}V3V_{3}V4V_{4}V5V_{5}
Figure 4.1. The situation in the proof of Theorem 4.1 in the basic case when Kj=ΩFjK_{j}=\Omega\cap\partial^{*}F_{j}. The essential partition of TT induced by KjT[sj]K_{j}\cup T[s_{j}] is denoted by {Uij}i\{U_{i}^{j}\}_{i}. The limit partition {Ui}i\{U_{i}\}_{i} of {Uij}i\{U_{i}^{j}\}_{i} may fail to be the essential partition of TT induced by K=TiUiK^{*}=T\cap\cup_{i}\partial^{*}U_{i}, since some of the UiU_{i} may be essentially disconnected. In the picture, denoting by {Vk}k\{V_{k}\}_{k} the essential partition of TT induced by KK^{*}, we have U5=V5V6=TFU_{5}=V_{5}\cup V_{6}=T\cap F. We also notice, in reference to the notation set in (4.6), that X1j={5}X_{1}^{j}=\{5\} and X0j={1,2,3,4}X_{0}^{j}=\{1,2,3,4\}.

Figure 4.1.

Proof of Theorem 4.1.

Step one: We start by showing that, for each jj and ii such that |Uij|>0|U_{i}^{j}|>0, we have

either(Uij)(1)Fj(1),or(Uij)(1)Fj(0),\displaystyle\mbox{either}\quad(U_{i}^{j})^{{\scriptscriptstyle{(1)}}}\subset F_{j}^{{\scriptscriptstyle{(1)}}}\,,\qquad\mbox{or}\quad(U_{i}^{j})^{{\scriptscriptstyle{(1)}}}\subset F_{j}^{\scriptscriptstyle{(0)}}\,, (4.4)

and for each ii such that |Ui|>0|U_{i}|>0,

eitherUi(1)F(1),orUi(1)F(0).\mbox{either}\quad U_{i}^{{\scriptscriptstyle{(1)}}}\subset F^{{\scriptscriptstyle{(1)}}}\,,\qquad\mbox{or}\quad U_{i}^{{\scriptscriptstyle{(1)}}}\subset F^{\scriptscriptstyle{(0)}}\,. (4.5)

Postponing for the moment the proof of (4.4) and (4.5), let us record several consequences of these inclusions. First, if we set

X1j={i:|Uij|>0,(Uij)(1)Fj(1)},X0j={i:|Uij|>0,(Uij)(1)Fj(0)},\displaystyle X^{j}_{1}=\big{\{}i:|U_{i}^{j}|>0\,,\,(U_{i}^{j})^{{\scriptscriptstyle{(1)}}}\subset F_{j}^{{\scriptscriptstyle{(1)}}}\big{\}}\,,\qquad X^{j}_{0}=\big{\{}i:|U_{i}^{j}|>0\,,\,(U_{i}^{j})^{{\scriptscriptstyle{(1)}}}\subset F_{j}^{\scriptscriptstyle{(0)}}\big{\}}\,, (4.6)
X1={i:|Ui|>0,Ui(1)F(1)},X0={i:|Ui|>0,Ui(1)F(0)},\displaystyle X_{1}=\big{\{}i:|U_{i}|>0\,,\,U_{i}^{{\scriptscriptstyle{(1)}}}\subset F^{{\scriptscriptstyle{(1)}}}\big{\}}\,,\qquad\hskip 14.22636ptX_{0}=\big{\{}i:|U_{i}|>0\,,\,U_{i}^{{\scriptscriptstyle{(1)}}}\subset F^{\scriptscriptstyle{(0)}}\big{\}}\,, (4.7)

then, thanks to (4.4) and (4.5), we have

Xj:={i:|Uij|>0}=X0jX1j,X:={i:|Ui|>0}=X0X1.\displaystyle X^{j}:=\{i:|U_{i}^{j}|>0\}=X_{0}^{j}\cup X_{1}^{j}\,,\qquad X:=\{i:|U_{i}|>0\}=X_{0}\cup X_{1}\,. (4.8)

Combining (4.4) and (4.5) with FjFF_{j}\to F and UijUiU_{i}^{j}\to U_{i}, we find that for every iXi\in X, there is JiJ_{i}\in\mathbb{N} such that, for every m{0,1}m\in\{0,1\},

if iXmi\in X_{m}, then iXmji\in X_{m}^{j} for all jJij\geq J_{i}. (4.9)

Lastly, {Uij}iX1j\{U_{i}^{j}\}_{i\in X_{1}^{j}} is a Lebesgue partition of TFjT\cap F_{j}, and thus, by Federer’s theorem (1.37),

TFj(1)niX1j(Uij)(1)Uij,TFjnTiX1jUijTKj.\displaystyle T\cap F_{j}^{\scriptscriptstyle{(1)}}\,\,\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\,\,\bigcup_{i\in X_{1}^{j}}(U_{i}^{j})^{\scriptscriptstyle{(1)}}\cup\partial^{*}U_{i}^{j}\,,\qquad T\cap\partial^{*}F_{j}\,\,\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\,\,T\cap\bigcup_{i\in X_{1}^{j}}\partial^{*}U_{i}^{j}\,\,\subset\,\,T\cap K_{j}^{*}\,. (4.10)

To prove (4.4) and (4.5): Since {Uij}i\{U_{i}^{j}\}_{i} is the essential partition of TT induced by KjT[sj]K_{j}\cup T[s_{j}] and Kj=UBEP(KjT[sj];T)K_{j}^{*}={\rm UBEP}(K_{j}\cup T[s_{j}];T), we have

Kj is n-contained in KjT[sj],j,\displaystyle\mbox{$K^{*}_{j}$ is $\mathcal{H}^{n}$-contained in $K_{j}\cup T[s_{j}]$}\,,\qquad\forall j\,, (4.11)
KjT[sj] does not essentially disconnect Uij,i,j.\displaystyle\mbox{$K_{j}\cup T[s_{j}]$ does not essentially disconnect $U_{i}^{j}$}\,,\qquad\forall i,j\,. (4.12)

Since ΩFj\Omega\cap\partial^{*}F_{j} is n\mathcal{H}^{n}-contained in KjT[sj]K_{j}\cup T[s_{j}], the combination of (4.12) with Federer’s theorem (1.37) gives (4.4). The combination of |UijΔUi|0|U_{i}^{j}\Delta U_{i}|\to 0 as jj\to\infty with (4.4) gives (4.5).

Step two: We reduce the proof of (4.2) to that of

n(Ui(1)T[s0])=0,iX0.\displaystyle\mathcal{H}^{n}(U_{i}^{\scriptscriptstyle{(1)}}\cap T[s_{0}])=0\,,\qquad\forall i\in X_{0}\,. (4.13)

Indeed, {Ui(1):iX0}{F(0)Ui:iX0}\{U_{i}^{\scriptscriptstyle{(1)}}:i\in X_{0}\}\cup\{F^{\scriptscriptstyle{(0)}}\cap\partial^{*}U_{i}:i\in X_{0}\} is an n\mathcal{H}^{n}-partition of TF(0)T\cap F^{\scriptscriptstyle{(0)}}. In particular, TF(0)T\cap F^{\scriptscriptstyle{(0)}} is n\mathcal{H}^{n}-contained in iX0Ui(1)Ui\cup_{i\in X_{0}}U_{i}^{{\scriptscriptstyle{(1)}}}\cup\partial^{*}U_{i}, so that, should (4.13) hold, then T[s0]F(0)T[s_{0}]\cap F^{\scriptscriptstyle{(0)}} would be n\mathcal{H}^{n}-contained in iX0Ui\cup_{i\in X_{0}}\partial^{*}U_{i}, and thus in KK^{*}, thus proving (4.2).

Step three: We change variables from TT to101010Here we identify SS1\SS^{1} with /(2π)\mathbb{R}/(2\pi\mathbb{Z}) and, with a slight abuse of notation, denote by n+1\mathcal{L}^{n+1} the “Lebesgue measure on SS1×B1n\SS^{1}\times B_{1}^{n}”, which we use to define sets of finite perimeter and points of density in SS1×B1n\SS^{1}\times B_{1}^{n}. Y=Φ1(T)=SS1×B1nY=\Phi^{-1}(T)=\SS^{1}\times B_{1}^{n}. We set Y[s]=Φ1(T[s])={s}×B1nY[s]=\Phi^{-1}(T[s])=\{s\}\times B_{1}^{n} for the ss-slice of YY, and

Yi=Φ1(Ui),Yij=Φ1(Uij),Wi=YYi,Wij=YYij,Y_{i}=\Phi^{-1}(U_{i})\,,\qquad Y_{i}^{j}=\Phi^{-1}(U_{i}^{j})\,,\qquad W_{i}=Y\setminus Y_{i}\,,\qquad W_{i}^{j}=Y\setminus Y_{i}^{j}\,, (4.14)

Since Φ\Phi is a diffeomorphism, by [KMS22a, Lemma A.1] and the area formula we have that

Φ1(H)=Φ1(H),(Φ1(H))(m)=Φ1(H(m)),m{0,1},\partial^{*}\Phi^{-1}(H)=\Phi^{-1}(\partial^{*}H)\,,\qquad(\Phi^{-1}(H))^{\scriptscriptstyle{(m)}}=\Phi^{-1}(H^{\scriptscriptstyle{(m)}})\,,m\in\{0,1\}\,, (4.15)

for every set of finite perimeter HTH\subset T; in particular, setting

Mj=Φ1(FjT),M=Φ1(FT),M_{j}=\Phi^{-1}(F_{j}\cap T)\,,\qquad M=\Phi^{-1}(F\cap T)\,,

by Federer’s theorem (1.37), we see that (4.1) is equivalent

Y[sj] is n-contained in iYijMj(1)Mj,\mbox{$Y[s_{j}]$ is $\mathcal{H}^{n}$-contained in $\bigcup_{i}\partial^{*}Y_{i}^{j}\cup M_{j}^{\scriptscriptstyle{(1)}}\cup\partial^{*}M_{j}$}\,, (4.16)

By (4.10) and (4.15), we may rewrite (4.16) as

Y[sj] is n-contained in iYijiX1j(Yij)(1).\mbox{$Y[s_{j}]$ is $\mathcal{H}^{n}$-contained in $\bigcup_{i\in\mathbb{N}}\partial^{*}Y_{i}^{j}\cup\bigcup_{i\in X_{1}^{j}}(Y_{i}^{j})^{\scriptscriptstyle{(1)}}$}\,. (4.17)

Similarly, Yi(1)=Φ1(Ui(1))Y_{i}^{\scriptscriptstyle{(1)}}=\Phi^{-1}(U_{i}^{\scriptscriptstyle{(1)}}) for every ii, and thus (4.13) is equivalent to

n(Yi(1)Y[s0])=0,iX0.\mathcal{H}^{n}(Y_{i}^{{\scriptscriptstyle{(1)}}}\cap Y[s_{0}])=0\,,\qquad\forall i\in X_{0}\,. (4.18)

We are thus left to prove that (4.17) implies (4.18).

To this end, let us denote by 𝐩\mathbf{p} the projection of Y=SS1×B1nY=\SS^{1}\times B_{1}^{n} onto B1nB_{1}^{n}, and consider the sets

Gi=𝐩(Yi(1)Y[s0]),Gi=GGi,G_{i}=\mathbf{p}\big{(}Y_{i}^{{\scriptscriptstyle{(1)}}}\cap Y[s_{0}]\big{)}\,,\qquad G_{i}^{*}=G^{*}\cap G_{i}\,,

corresponding to the set GB1nG^{*}\subset B_{1}^{n} with n(B1nG)=0\mathcal{H}^{n}(B_{1}^{n}\setminus G^{*})=0 defined as follows:

(i) denoting by Hy={s𝕊1:(s,y)H}H_{y}=\{s\in\mathbb{S}^{1}:(s,y)\in H\} the “circular slice of HYH\subset Y above yy”, if yGy\in G^{*}, jj\in\mathbb{N}, kk is an index for the partitions {Yk}k\{Y_{k}\}_{k} and {Ykj}\{Y_{k}^{j}\}, and H{Yk,Wk,Ykj,Wkj}H\in\{Y_{k},W_{k},Y_{k}^{j},W_{k}^{j}\}, then HyH_{y} is a set of finite perimeter in 𝕊1\mathbb{S}^{1} with

Hy=1(Hy)(1)𝕊1,𝕊1(Hy)=0(H)y,H_{y}\overset{\mathcal{H}^{1}}{=}(H_{y})^{{{\scriptscriptstyle{(1)}}}_{\mathbb{S}^{1}}}\,,\qquad\partial^{*}_{\mathbb{S}^{1}}(H_{y})\overset{\mathcal{H}^{0}}{=}(\partial^{*}H)_{y}\,, (4.19)

(and thus with 𝕊1(Hy)=(H)y\partial^{*}_{\mathbb{S}^{1}}(H_{y})=(\partial^{*}H)_{y}); this is a standard consequence of the slicing theory for sets of finite perimeter, see, e.g., [BCF13, Theorem 2.4] or [Mag12, Remark 18.13];

(ii) for every yGy\in G^{*} and jj\in\mathbb{N},

(sj,y)kYkjkX1j(Ykj)(1);(s_{j},y)\in\bigcup_{k\in\mathbb{N}}\partial^{*}Y_{k}^{j}\cup\bigcup_{k\in X_{1}^{j}}(Y_{k}^{j})^{\scriptscriptstyle{(1)}}\,; (4.20)

this is immediate from (4.17);

(iii) for every yGy\in G^{*}, and kk an index for the partitions {Yk}k\{Y_{k}\}_{k} and {Ykj}\{Y_{k}^{j}\},

limj1((Yk)yΔ(Ykj)y)=0;\lim_{j\to\infty}\mathcal{H}^{1}((Y_{k})_{y}\Delta(Y_{k}^{j})_{y})=0\,; (4.21)

this is immediate from Fubini’s theorem and YkjYkY_{k}^{j}\to Y_{k} as jj\to\infty;

(iv) for every yGy\in G^{*},

k0((Ykj)y)<;\displaystyle\sum_{k}\mathcal{H}^{0}((\partial^{*}Y_{k}^{j})_{y})<\infty; (4.22)

indeed, by applying in the order the coarea formula, the area formula and (2.3) we find

kB1n0((Ykj)y)𝑑n\displaystyle\sum_{k}\int_{B_{1}^{n}}\mathcal{H}^{0}((\partial^{*}Y_{k}^{j})_{y})\,d\mathcal{H}^{n} \displaystyle\leq kP(Ykj;Y)(LipΦ1)nkP(Ukj;T)\displaystyle\sum_{k}P(Y_{k}^{j};Y)\leq({\rm Lip}\Phi^{-1})^{n}\,\sum_{k}P(U_{k}^{j};T)
\displaystyle\leq 2(LipΦ1)nn(KjT[sj]).\displaystyle 2\,({\rm Lip}\Phi^{-1})^{n}\,\mathcal{H}^{n}(K_{j}\cup T[s_{j}])\,.

Now, let us pick yGiy\in G_{i}^{*}. Since yGiy\in G_{i} implies (s0,y)Yi(1)(s_{0},y)\in Y_{i}^{\scriptscriptstyle{(1)}}, and Yi(1)Yi=Y_{i}^{\scriptscriptstyle{(1)}}\cap\partial^{*}Y_{i}=\varnothing, we find (s0,y)Yi(s_{0},y)\not\in\partial^{*}Y_{i}, i.e. s0(Yi)ys_{0}\not\in(\partial^{*}Y_{i})_{y}. By yGy\in G^{*}, we have (Yi)y=𝕊1(Yi)y(\partial^{*}Y_{i})_{y}=\partial^{*}_{\mathbb{S}^{1}}(Y_{i})_{y}, so that

s0𝕊1(Yi)y.s_{0}\not\in\partial^{*}_{\mathbb{S}^{1}}(Y_{i})_{y}\,. (4.23)

Since (Yi)y(Y_{i})_{y} has finite perimeter, 𝕊1(Yi)y\partial^{*}_{\mathbb{S}^{1}}(Y_{i})_{y} is a finite set, and so (4.23) implies the existence of an open interval 𝒜y𝕊1\mathcal{A}_{y}\subset\mathbb{S}^{1}, containing s0s_{0}, 1\mathcal{H}^{1}-contained either in (Yi)y(Y_{i})_{y} or in (Wi)y(W_{i})_{y}, and such that

𝕊1𝒜y(Yi)y=SS1(Wi)y.\partial_{\mathbb{S}^{1}}\mathcal{A}_{y}\subset(\partial^{*}Y_{i})_{y}=\partial^{*}_{\SS^{1}}(W_{i})_{y}\,. (4.24)

We claim that there is GiGiG_{i}^{**}\subset G_{i}^{*}, with full n\mathcal{H}^{n}-measure in GiG_{i}^{*} (and thus in GiG_{i}), such that

𝒜y is 1-contained in (Yi)y,yGi.\mbox{$\mathcal{A}_{y}$ is $\mathcal{H}^{1}$-contained in $(Y_{i})_{y}$}\,,\qquad\forall y\in G_{i}^{**}\,. (4.25)

Indeed, let us consider the countable decomposition {Gi,m}m=1\{G_{i,m}^{*}\}_{m=1}^{\infty} of GiG_{i}^{*} given by

Gi,m={yGi:dist({s0},SS1𝒜y)[1/(m+1),1/m)}B1n,G_{i,m}^{*}=\Big{\{}y\in G_{i}^{*}:{\rm dist}\big{(}\{s_{0}\},\partial_{\SS^{1}}\mathcal{A}_{y}\big{)}\in\big{[}1\big{/}(m+1),1\big{/}m\big{)}\Big{\}}\subset B_{1}^{n}\,,

and let

Zi,m={yGi,m:𝒜y is 1-contained in (Wi)y}.Z_{i,m}=\big{\{}y\in G_{i,m}^{*}:\mbox{$\mathcal{A}_{y}$ is $\mathcal{H}^{1}$-contained in $(W_{i})_{y}$}\big{\}}\,.

If n(Zi,m)>0\mathcal{H}^{n}(Z_{i,m})>0, then there is yZi,m(1)y^{*}\in Z_{i,m}^{\scriptscriptstyle{(1)}}, so that n(Zi,mBrn(y))=ωnrn+o(rn)\mathcal{H}^{n}(Z_{i,m}\cap B_{r}^{n}(y^{*}))=\omega_{n}\,r^{n}+{\rm o}(r^{n}). Therefore, if r<1/(m+1)r<1/(m+1) and Br1(s0)B_{r}^{1}(s_{0}) denotes the open interval of center s0s_{0} and radius rr inside SS1\SS^{1}, then

n+1(Yi(Br1(s0)×Brn(y)))=Brn(y)1(Br1(s0)(Yi)y)𝑑yn\displaystyle\mathcal{L}^{n+1}\big{(}Y_{i}\cap\big{(}B_{r}^{1}(s_{0})\times B_{r}^{n}(y^{*})\big{)}\big{)}=\int_{B_{r}^{n}(y^{*})}\mathcal{H}^{1}(B_{r}^{1}(s_{0})\cap(Y_{i})_{y})\,d\mathcal{H}^{n}_{y}
=\displaystyle= Zi,mBrn(y)1(Br1(s0)(Yi)y)𝑑yn+o(rn+1)=o(rn+1)\displaystyle\int_{Z_{i,m}\cap B_{r}^{n}(y^{*})}\mathcal{H}^{1}(B_{r}^{1}(s_{0})\cap(Y_{i})_{y})\,d\mathcal{H}^{n}_{y}+{\rm o}(r^{n+1})={\rm o}(r^{n+1})

where in the last identity we have used the facts that yZi,mBrn(y)y\in Z_{i,m}\cap B_{r}^{n}(y^{*}), s0𝒜ys_{0}\in\mathcal{A}_{y}, and r<1/(m+1)r<1/(m+1) to conclude that Br1(s0)B_{r}^{1}(s_{0}) is 1\mathcal{H}^{1}-contained in (Wi)y(W_{i})_{y}; in particular, (s0,y)Yi(0)(s_{0},y^{*})\in Y_{i}^{\scriptscriptstyle{(0)}}, against the fact that Zi,mGi(=𝐩(Y[s0]Yi(1)))Z_{i,m}\subset G_{i}(=\mathbf{p}(Y[s_{0}]\cap Y_{i}^{\scriptscriptstyle{(1)}})). We have thus proved that each Zi,mZ_{i,m} is n\mathcal{H}^{n}-negligible, and therefore that there is GiGiG_{i}^{**}\subset G_{i}^{*} and n\mathcal{H}^{n}-equivalent to GiG_{i}^{*}, such that (4.25) holds true.

Having proved (4.25), we now notice that, by (4.20), yGiy\in G_{i}^{*} implies

sjk(Ykj)ykX1j((Ykj)(1))y=kSS1(Ykj)ykX1j((Ykj)y)(1)𝕊1.s_{j}\in\bigcup_{k\in\mathbb{N}}(\partial^{*}Y_{k}^{j})_{y}\cup\bigcup_{k\in X_{1}^{j}}\big{(}(Y_{k}^{j})^{\scriptscriptstyle{(1)}}\big{)}_{y}=\bigcup_{k}\partial^{*}_{\SS^{1}}(Y_{k}^{j})_{y}\cup\bigcup_{k\in X_{1}^{j}}\big{(}(Y_{k}^{j})_{y}\big{)}^{{\scriptscriptstyle{(1)}}_{\mathbb{S}^{1}}}\,. (4.26)

If (4.26) holds because sjSS1(Ykj)ys_{j}\in\partial^{*}_{\SS^{1}}(Y_{k}^{j})_{y} for some kk, then, thanks to (4.22) there must kkk^{\prime}\neq k such that sjSS1(Ykj)ys_{j}\in\partial^{*}_{\SS^{1}}(Y_{k^{\prime}}^{j})_{y} too; since either kk or kk^{\prime} must be different from ii, we conclude that siSS1(Yk(i)j)ys_{i}\in\partial^{*}_{\SS^{1}}(Y_{k(i)}^{j})_{y} for some k(i)ik(i)\neq i; if, instead, (4.26) holds because sj((Ykj)y)(1)𝕊1s_{j}\in\big{(}(Y_{k}^{j})_{y}\big{)}^{{\scriptscriptstyle{(1)}}_{\mathbb{S}^{1}}} for some kX1jk\in X_{1}^{j}, then we can recall that, thanks to (4.9), iX0ji\in X_{0}^{j} for every jJij\geq J_{i}, and thus iki\neq k; in summary, for each yGiy\in G_{i}^{*},

if jJi, then k(j)i s.t.sj𝕊1(Yk(j)j)y((Yk(j)j)y)(1)𝕊1.\mbox{if $j\geq J_{i}$, then $\exists k(j)\neq i$ s.t.}\,\,s_{j}\in\partial^{*}_{\mathbb{S}^{1}}(Y_{k(j)}^{j})_{y}\cup\big{(}(Y_{k(j)}^{j})_{y}\big{)}^{{\scriptscriptstyle{(1)}}_{\mathbb{S}^{1}}}\,. (4.27)

With the goal of obtaining a lower bound on the relative perimeters of the sets YijY_{i}^{j} in a neighborhood of GiG_{i} (see (4.31) below), we now consider yGiy\in G_{i}^{**}, and pick r>0r>0 such that clBr1(s0)𝒜y\mathrm{cl}\,B_{r}^{1}(s_{0})\subset\mathcal{A}_{y}. Correspondingly, since sjs0s_{j}\to s_{0} and (4.27) holds, we can find J=J(i,y,r)JiJ^{*}=J^{*}(i,y,r)\geq J_{i} such that, for jJj\geq J^{*},

sjBr1(s0)[𝕊1(Yk(j)j)y((Yk(j)j)y)(1)𝕊1]𝒜y[𝕊1(Yk(j)j)y((Yk(j)j)y)(1)𝕊1].s_{j}\in B_{r}^{1}(s_{0})\cap\big{[}\partial^{*}_{\mathbb{S}^{1}}(Y_{k(j)}^{j})_{y}\cup\big{(}(Y_{k(j)}^{j})_{y}\big{)}^{{\scriptscriptstyle{(1)}}_{\mathbb{S}^{1}}}\big{]}\subset\mathcal{A}_{y}\cap\big{[}\partial^{*}_{\mathbb{S}^{1}}(Y_{k(j)}^{j})_{y}\cup\big{(}(Y_{k(j)}^{j})_{y}\big{)}^{{\scriptscriptstyle{(1)}}_{\mathbb{S}^{1}}}\big{]}\,. (4.28)

Now, by (4.21), k(j)ik(j)\neq i, and 𝒜y1(Yi)y\mathcal{A}_{y}\overset{\mathcal{H}^{1}}{\subset}(Y_{i})_{y}, we have

limj1(𝒜y(Yk(j)j)y)=0.\lim_{j\to\infty}\mathcal{H}^{1}(\mathcal{A}_{y}\cap(Y_{k(j)}^{j})_{y})=0\,. (4.29)

Since, by (4.19), (Yk(j)j)y(Y_{k(j)}^{j})_{y} is 1\mathcal{H}^{1}-equivalent to a finite union of intervals, (4.28) implies the existence of an open interval yj\mathcal{I}_{y}^{j} such that

sjclSS1yj,yj1(Yk(j)j)y,SS1yj(Yk(j)j)y(Wij)y,\displaystyle s_{j}\in\mathrm{cl}\,_{\SS^{1}}\mathcal{I}_{y}^{j}\,,\qquad\mathcal{I}_{y}^{j}\overset{\mathcal{H}^{1}}{\subset}(Y_{k(j)}^{j})_{y}\,,\qquad\partial_{\SS^{1}}\mathcal{I}_{y}^{j}\subset(\partial^{*}Y_{k(j)}^{j})_{y}\subset(\partial^{*}W_{i}^{j})_{y}\,, (4.30)

which, due to (4.28) and (4.29), must satisfy

limjdiam(yj)=0.\lim_{j\to\infty}{\rm diam}\,\big{(}\mathcal{I}_{y}^{j}\big{)}=0\,.

In particular,

SS1yjBr1(s0),jJ,\partial_{\SS^{1}}\,\mathcal{I}_{y}^{j}\subset B_{r}^{1}(s_{0})\,,\qquad\forall j\geq J^{*}\,,

and thus, by the last inclusion in (4.30),

0(Br1(s0)SS1(Wij)y)0(Br1(s0)SS1jy)2,\mathcal{H}^{0}\big{(}B_{r}^{1}(s_{0})\cap\partial^{*}_{\SS^{1}}(W_{i}^{j})_{y}\big{)}\geq\mathcal{H}^{0}(B_{r}^{1}(s_{0})\cap\partial_{\SS^{1}}\mathcal{I}_{j}^{y})\geq 2\,,

whenever jJj\geq J^{*}. Since yGiy\in G_{i}^{**} and r>0r>0 were arbitrary, by the coarea formula and Fatou’s lemma,

lim infjP(Wij;Br1(s0)×Gi)\displaystyle\liminf_{j\to\infty}P(W_{i}^{j};B_{r}^{1}(s_{0})\times G_{i}^{**}) \displaystyle\geq lim infjGi0(Br1(s0)SS1(Wij)y)𝑑yn\displaystyle\liminf_{j\to\infty}\int_{G_{i}^{**}}\mathcal{H}^{0}\big{(}B_{r}^{1}(s_{0})\cap\partial^{*}_{\SS^{1}}(W_{i}^{j})_{y}\big{)}\,d\mathcal{H}^{n}_{y} (4.31)
\displaystyle\geq 2n(Gi)=2n(Gi).\displaystyle 2\,\mathcal{H}^{n}(G_{i}^{**})=2\,\mathcal{H}^{n}(G_{i})\,.

Now, since Wij=Yij=Φ1(Uij)\partial^{*}W_{i}^{j}=\partial^{*}Y_{i}^{j}=\Phi^{-1}(\partial^{*}U_{i}^{j}), by (4.11) we have

YiWij is n-contained in Y[sj]Φ1(TKj),\mbox{$Y\cap\bigcup_{i}\partial^{*}W_{i}^{j}$ is $\mathcal{H}^{n}$-contained in $Y[s_{j}]\cup\Phi^{-1}\big{(}T\cap K_{j}\big{)}$}\,,

which implies, for every jj large enough to have sjBr1(s0)s_{j}\in B_{r}^{1}(s_{0}),

P(Wij;Br1(s0)×Gi)\displaystyle P(W_{i}^{j};B_{r}^{1}(s_{0})\times G_{i}^{**})
n(Y[sj](Br1(s0)×G1))+n(Φ1(TKj)(Br1(s0)×B1n))\displaystyle\leq\mathcal{H}^{n}\big{(}Y[s_{j}]\cap(B_{r}^{1}(s_{0})\times G_{1}^{**})\big{)}+\mathcal{H}^{n}\big{(}\Phi^{-1}(T\cap K_{j})\cap(B_{r}^{1}(s_{0})\times B_{1}^{n})\big{)}
=n(Gi)+n(Φ1(TKj)(Br1(s0)×B1n))\displaystyle=\mathcal{H}^{n}(G_{i}^{**})+\mathcal{H}^{n}\big{(}\Phi^{-1}(T\cap K_{j})\cap(B_{r}^{1}(s_{0})\times B_{1}^{n})\big{)}
n(Gi)+Lip(Φ1)nn(KjΦ(Br1(s0)×B1n)).\displaystyle\leq\mathcal{H}^{n}(G_{i})+{\rm Lip}(\Phi^{-1})^{n}\,\mathcal{H}^{n}\big{(}K_{j}\cap\Phi(B_{r}^{1}(s_{0})\times B_{1}^{n})\big{)}\,. (4.32)

By combining (4.31) with (4.32) we conclude that for every r>0r>0

n(Gi)Lip(Φ1)nμ(Φ(cl(Br1(s0))×B1n)),\mathcal{H}^{n}(G_{i})\leq{\rm Lip}(\Phi^{-1})^{n}\,\mu\big{(}\Phi(\mathrm{cl}\,(B_{r}^{1}(s_{0}))\times B_{1}^{n})\big{)}\,, (4.33)

By μ(T[s0])=0\mu(T[s_{0}])=0, if we let r0+r\to 0^{+} in (4.33), we conclude that n(Gi)=0\mathcal{H}^{n}(G_{i})=0. Now, since Gi=𝐩(Yi(1)Y[s0])G_{i}=\mathbf{p}\big{(}Y_{i}^{{\scriptscriptstyle{(1)}}}\cap Y[s_{0}]\big{)}, we have

n(Yi(1)Y[s0])=n(Gi),\mathcal{H}^{n}\big{(}Y_{i}^{{\scriptscriptstyle{(1)}}}\cap Y[s_{0}]\big{)}=\mathcal{H}^{n}(G_{i})\,, (4.34)

thus proving (4.18), and hence the theorem. ∎

5. Direct Method on generalized soap films (Theorem 1.4)

In Section 5.1 we prove Theorem 1.4, while in Section 5.2 we notice the changes to that argument that are needed to prove a different closure theorem that will be crucial in the companion papers [MNR23a, MNR23b]. In particular, Section 5.2 will not be needed for the other main results of this paper (although it is included here since it is definitely easier to understand in this context).

5.1. Proof of Theorem 1.4

Let us first of all recall the setting of the theorem. We are given a closed set 𝐖\mathbf{W} in n+1\mathbb{R}^{n+1}, a spanning class 𝒞\mathcal{C} for 𝐖\mathbf{W}, and a sequence {(Kj,Ej)}j\{(K_{j},E_{j})\}_{j} in 𝒦B\mathcal{K}_{\rm B} such that

supjn(Kj)<,\sup_{j}\,\mathcal{H}^{n}(K_{j})<\infty\,, (5.1)

and, for some Borel set EE and Radon measures μbk\mu_{\rm bk} and μbd\mu_{\rm bd} in Ω\Omega, it holds that EjlocEE_{j}\stackrel{{\scriptstyle\scriptscriptstyle{{\rm loc}}}}{{\to}}E and

n (ΩEj)+2n ((Kj)Ej(0))μbk,\displaystyle\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E_{j})+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(0)}})\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}\mu_{\rm bk}\,, (5.2)
n (ΩEj)+2n ((Kj)Ej)μbd,\displaystyle\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E_{j})+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\mathcal{R}(K_{j})\setminus\partial^{*}E_{j})\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}\mu_{\rm bd}\,, (5.3)

as jj\to\infty. In this setting we want to prove that the sets

Kbk\displaystyle K_{\rm bk}\!\! :=\displaystyle:= (ΩE){xΩE(0):θn(μbk)(x)2},\displaystyle\!\!\big{(}\Omega\cap\partial^{*}E\big{)}\cup\Big{\{}x\in\Omega\cap E^{\scriptscriptstyle{(0)}}:\theta^{n}_{*}(\mu_{\rm bk})(x)\geq 2\Big{\}}\,, (5.4)
Kbd\displaystyle K_{\rm bd}\!\! :=\displaystyle:= (ΩE){xΩE:θn(μbd)(x)2},\displaystyle\!\!\big{(}\Omega\cap\partial^{*}E\big{)}\cup\Big{\{}x\in\Omega\setminus\partial^{*}E:\theta^{n}_{*}(\mu_{\rm bd})(x)\geq 2\Big{\}}\,, (5.5)

are such that (Kbk,E),(Kbd,E)𝒦B(K_{\rm bk},E),(K_{\rm bd},E)\in\mathcal{K}_{\rm B} and

μbk\displaystyle\mu_{\rm bk}\!\! \displaystyle\geq n (ΩE)+2n (KbkE(0)),\displaystyle\!\!\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E)+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{\rm bk}\cap E^{\scriptscriptstyle{(0)}})\,, (5.6)
μbd\displaystyle\mu_{\rm bd}\!\! \displaystyle\geq n (ΩE)+2n (KbdE),\displaystyle\!\!\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E)+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{\rm bd}\setminus\partial^{*}E)\,, (5.7)

with

lim infjbk(Kj,Ej)bk(Kbk,E),lim infjbd(Kj,Ej)bd(Kbd,E);\liminf_{j\to\infty}\mathcal{F}_{\rm bk}(K_{j},E_{j})\geq\mathcal{F}_{\rm bk}(K_{\rm bk},E)\,,\qquad\liminf_{j\to\infty}\mathcal{F}_{\rm bd}(K_{j},E_{j})\geq\mathcal{F}_{\rm bd}(K_{\rm bd},E)\,; (5.8)

and that the closure statements

if KjEj(1)K_{j}\cup E_{j}^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} for every jj, (5.9)
then KbkE(1) is 𝒞-spanning 𝐖,\displaystyle\mbox{then $K_{\rm bk}\cup E^{\scriptscriptstyle{(1)}}$ is $\mathcal{C}$-spanning $\mathbf{W}$}\,, (5.10)

and

if KjK_{j} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} for every jj, (5.11)
then Kbd is 𝒞-spanning 𝐖,\displaystyle\mbox{then $K_{\rm bd}$ is $\mathcal{C}$-spanning $\mathbf{W}$}\,, (5.12)

hold true.

Proof of Theorem 1.4.

By ΩEKbkKbd\Omega\cap\partial^{*}E\subset K_{\rm bk}\cap K_{\rm bd} we have (Kbk,E),(Kbd,E)𝒦B(K_{\rm bk},E),(K_{\rm bd},E)\in\mathcal{K}_{\rm B}. By [Mag12, Theorem 6.4], θn(μbk)2\theta^{n}_{*}(\mu_{\rm bk})\geq 2 on KbkE(0)K_{\rm bk}\cap E^{\scriptscriptstyle{(0)}} implies μbk  (KbkE(0))2n  (KbkE(0))\mu_{\rm bk}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{\rm bk}\cap E^{\scriptscriptstyle{(0)}})\geq 2\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{\rm bk}\cap E^{\scriptscriptstyle{(0)}}), and, similarly, we have μbd  (KbdE)2n  (KbdE)\mu_{\rm bd}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{\rm bd}\setminus\partial^{*}E)\geq 2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{\rm bd}\setminus\partial^{*}E). Since, by the lower semicontinuity of distributional perimeter, we have min{μbk,μbd}n  (EΩ)\min\{\mu_{\rm bk},\mu_{\rm bd}\}\geq\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\partial^{*}E\cap\Omega), (5.6), (5.7) and (5.8) follow. We are thus left to prove that if either (5.9) or (5.11) holds, then (5.10) or (5.12) holds respectively. We divide the proof into three parts, numbered by Roman numerals.

I. Set up of the proof: Fixing from now on a choice of (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}) against which we want to test the 𝒞\mathcal{C}-spanning properties (5.10) and (5.12), we introducing several key objects related to (γ,Φ,T)(\gamma,\Phi,T).

Introducing s0s_{0}: Up to extracting subsequences, let μ\mu be the weak-star limit of n  Kj\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}\,K_{j}, and set

J={sSS1:μ(T[s])=0},\displaystyle J=\{s\in\SS^{1}:\mu(T[s])=0\}\,, (5.13)

so that 1(SS1J)=0\mathcal{H}^{1}(\SS^{1}\setminus J)=0. We fix s0Js_{0}\in J.

Introducing sjs_{j}, {Uij}i\{U_{i}^{j}\}_{i}, and KjK_{j}^{*}: For 1\mathcal{H}^{1}-a.e. sSS1s\in\SS^{1} it holds that n(KjT[s])=0\mathcal{H}^{n}(K_{j}\cap T[s])=0 for every jj and (thanks to Theorem 1.3/Theorem 3.1) the essential partition {Uij[s]}i\{U_{i}^{j}[s]\}_{i} induced on TT by KjT[s]K_{j}\cup T[s] is such that

T[s]Ej(0) is n-contained in UBEP(KjT[s];T),\displaystyle\mbox{$T[s]\cap E_{j}^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in ${\rm UBEP}(K_{j}\cup T[s];T)$}\,, (if (5.9) holds),\displaystyle\qquad\mbox{(if \eqref{hp bulk spanning} holds)}\,,
T[s] is n-contained in UBEP(KjT[s];T),\displaystyle\mbox{$T[s]$ is $\mathcal{H}^{n}$-contained in ${\rm UBEP}(K_{j}\cup T[s];T)$}\,, (if (5.11) holds).\displaystyle\qquad\mbox{(if \eqref{hp interface spanning} holds)}\,.

Therefore we can find a sequence sjs0s_{j}\to s_{0} as jj\to\infty such that

n(KjT[sj])=0j,\mathcal{H}^{n}(K_{j}\cap T[s_{j}])=0\qquad\forall j\,, (5.14)

and, denoting by {Uij}i\{U_{i}^{j}\}_{i} the essential partition of TT induced by KjT[sj]K_{j}\cup T[s_{j}] (i.e. Uij=Uij[sj]U_{i}^{j}=U_{i}^{j}[s_{j}]), and setting for brevity

Kj=UBEP(KjT[sj];T)=TiUij,K_{j}^{*}={\rm UBEP}(K_{j}\cup T[s_{j}];T)=T\cap\bigcup_{i}\partial^{*}U_{i}^{j}\,, (5.15)

we have

T[sj]Ej(0) is n-contained in Kj,\displaystyle\mbox{$T[s_{j}]\cap E_{j}^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in $K_{j}^{*}$}\,, (if (5.9) holds),\displaystyle\qquad\mbox{(if \eqref{hp bulk spanning} holds)}\,, (5.16)
T[sj] is n-contained in Kj,\displaystyle\mbox{$T[s_{j}]$ is $\mathcal{H}^{n}$-contained in $K_{j}^{*}$}\,, (if (5.11) holds).\displaystyle\qquad\mbox{(if \eqref{hp interface spanning} holds)}\,. (5.17)

Introducing {Ui}i\{U_{i}\}_{i} and KK^{*}: By (5.1), Lemma 2.3, and up to extract a subsequence we can find a Lebesgue partition {Ui}i\{U_{i}\}_{i} of TT such that,

{Ui}i is the limit of {{Uij}i}j in the sense specified by (2.8).\mbox{$\{U_{i}\}_{i}$ is the limit of $\{\{U_{i}^{j}\}_{i}\}_{j}$ in the sense specified by \eqref{partition convergence}}\,. (5.18)

Correspondingly we set

K=TiUi.K^{*}=T\cap\bigcup_{i}\partial^{*}U_{i}\,. (5.19)

Having introduced s0s_{0}, sjs_{j}, {Uij}i\{U_{i}^{j}\}_{i}, KjK_{j}^{*}, {Ui}i\{U_{i}\}_{i}, and KK^{*}, we notice that if (5.9) holds, then we can apply Theorem 4.1 with Fj=EjF_{j}=E_{j} and find that

T[s0]E(0) is n-contained in K,\displaystyle\mbox{$T[s_{0}]\cap E^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in $K^{*}$}\,, (if (5.9) holds);\displaystyle\mbox{(if \eqref{hp bulk spanning} holds)}\,; (5.20)

if, instead, (5.11) holds, then Theorem 4.1 can be applied with Fj=F=F_{j}=F=\varnothing to deduce

T[s0] is n-contained in K,\displaystyle\mbox{$T[s_{0}]$ is $\mathcal{H}^{n}$-contained in $K^{*}$}\,, (if (5.11) holds).\displaystyle\mbox{(if \eqref{hp interface spanning} holds)}\,. (5.21)

We now make the following claim:

Claim: We have

K(T[s0]E(1)) is n-contained in Kbk,\displaystyle\mbox{$K^{*}\setminus(T[s_{0}]\cup E^{\scriptscriptstyle{(1)}})$ is $\mathcal{H}^{n}$-contained in $K_{\rm bk}$}\,, (5.22)
KT[s0] is n-contained in Kbd.\displaystyle\mbox{$K^{*}\setminus T[s_{0}]$ is $\mathcal{H}^{n}$-contained in $K_{\rm bd}$}\,. (5.23)

The rest of the proof of the theorem is then divided in two parts: the conclusion follows from the claim, and the proof of the claim.

II. Conclusion of the proof from the claim: Proof that (5.11) implies (5.12): By 1(SS1J)=0\mathcal{H}^{1}(\SS^{1}\setminus J)=0, the arbitrariness of s0Js_{0}\in J, and that of (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}), thanks to Theorem 1.3 we can conclude that KbdK_{\rm bd} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} by showing that

T[s0]T[s_{0}] is n\mathcal{H}^{n}-contained in UBEP(KbdT[s0];T){\rm UBEP}(K_{\rm bd}\cup T[s_{0}];T). (5.24)

Now, since {Ui}i\{U_{i}\}_{i} is a Lebesgue partition of TT induced by KK^{*} (in the very tautological sense that KK^{*} is defined as TiUiT\cap\cup_{i}\partial^{*}U_{i}!) and, by (5.23) in claim one, KK^{*} is n\mathcal{H}^{n}-contained in KbdT[s0]K_{\rm bd}\cup T[s_{0}], by Theorem 2.1-(a) we have that if {Zi}i\{Z_{i}\}_{i} is the essential partition of TT induced by KbdT[s0]K_{\rm bd}\cup T[s_{0}], then iUi\cup_{i}\partial^{*}U_{i} is n\mathcal{H}^{n}-contained in iZi\cup_{i}\partial^{*}Z_{i}: therefore, by definition of KK^{*} and by definition of UBEP{\rm UBEP}, we have that

K is n-contained in UBEP(KbdT[s0];T).\mbox{$K^{*}$ is $\mathcal{H}^{n}$-contained in ${\rm UBEP}\big{(}K_{\rm bd}\cup T[s_{0}];T\big{)}$}\,. (5.25)

By combining (5.25) with (5.21) we immediately deduce (5.24) and conclude.

Proof that (5.9) implies (5.10): Thanks to Theorem 3.1 it suffices to prove that

T[s0]E(0) is n-contained in UBEP(KbkT[s0];T).\mbox{$T[s_{0}]\cap E^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in ${\rm UBEP}(K_{\rm bk}\cup T[s_{0}];T)$}\,. (5.26)

By (5.20), the proof of (5.26) can be reduced to that of

KE(0) is n-contained in UBEP(KbkT[s0];T).\mbox{$K^{*}\cap E^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in ${\rm UBEP}(K_{\rm bk}\cup T[s_{0}];T)$}\,. (5.27)

Now, let us consider the Lebesgue partition of TT defined by {Vk}k={UiE}i{TE}\{V_{k}\}_{k}=\{U_{i}\setminus E\}_{i}\cup\{T\cap E\}. By [Mag12, Theorem 16.3] we easily see that for each ii

E(0)Uin(UiE)n(E(0)Ui)E,E^{\scriptscriptstyle{(0)}}\cap\partial^{*}U_{i}\,\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\,\partial^{*}(U_{i}\setminus E)\,\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\,\big{(}E^{\scriptscriptstyle{(0)}}\cap\partial^{*}U_{i}\big{)}\cup\partial^{*}E\,, (5.28)

which combined with T(TE)=TEKbkT\cap\partial^{*}(T\cap E)=T\cap\partial^{*}E\subset K_{\rm bk} and with (5.22) in claim one, gives

TkVk\displaystyle T\cap\bigcup_{k}\partial^{*}V_{k} =\displaystyle= (TE){Ti(UiE)}n(TE)(E(0)K)\displaystyle(T\cap\partial^{*}E)\cup\Big{\{}T\cap\bigcup_{i}\partial^{*}(U_{i}\setminus E)\Big{\}}\,\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\,(T\cap\partial^{*}E)\cup\,\big{(}E^{\scriptscriptstyle{(0)}}\cap K^{*}\big{)} (5.29)
n\displaystyle\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset} (TE)(KE(1))nKbkT[s0].\displaystyle(T\cap\partial^{*}E)\cup\,\big{(}K^{*}\setminus E^{\scriptscriptstyle{(1)}})\,\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\,K_{\rm bk}\cup T[s_{0}]\,.

By (5.29) we can exploit Theorem 2.1-(a) to conclude that

TkVk is n-contained in UBEP(KbkT[s0];T).\mbox{$T\cap\bigcup_{k}\partial^{*}V_{k}$ is $\mathcal{H}^{n}$-contained in ${\rm UBEP}(K_{\rm bk}\cup T[s_{0}];T)$}\,. (5.30)

By the first inclusion in (5.28), E(0)KE^{\scriptscriptstyle{(0)}}\cap K^{*} is n\mathcal{H}^{n}-contained in TkVkT\cap\bigcup_{k}\partial^{*}V_{k}, therefore (5.30) implies (5.27), as required. We are thus left to prove the two claims.

III. Proof of the claim: We finally prove that K(T[s0]E(1))K^{*}\setminus(T[s_{0}]\cup E^{\scriptscriptstyle{(1)}}) is n\mathcal{H}^{n}-contained in KbkK_{\rm bk} (that is (5.22)), and that KT[s0]K^{*}\setminus T[s_{0}] is n\mathcal{H}^{n}-contained in KbdK_{\rm bd} (that is (5.23)).

To this end, repeating the argument in the proof of Theorem 4.1 with Fj=EjF_{j}=E_{j} and F=EF=E we see that, if we set Xmj={i:(Uij)(1)Ej(m)}X^{j}_{m}=\{i:(U_{i}^{j})^{\scriptscriptstyle{(1)}}\subset E_{j}^{\scriptscriptstyle{(m)}}\} and Xm={i:Ui(1)E(m)}X_{m}=\{i:U_{i}^{\scriptscriptstyle{(1)}}\subset E^{\scriptscriptstyle{(m)}}\} for m{0,1}m\in\{0,1\} (see (4.6) and (4.7)), then

Xj:={i:|Uij|>0}=X0jX1j,X:={i:|Ui|>0}=X0X1;\displaystyle X^{j}:=\{i:|U_{i}^{j}|>0\}=X_{0}^{j}\cup X_{1}^{j}\,,\qquad X:=\{i:|U_{i}|>0\}=X_{0}\cup X_{1}\,; (5.31)

and, moreover, for every ii there is j(i)j(i) such that iXmi\in X_{m} implies iXmji\in X_{m}^{j} for every jj(i)j\geq j(i). Thanks to (5.31) we easily see that Kj=TiUijK_{j}^{*}=T\cap\cup_{i}\partial^{*}U_{i}^{j} can be decomposed as

Kj=n(i,k)X0j×X0j,ijMikj(i,k)X1j×X1j,ijMikj(i,k)X0j×X1jMikj,K_{j}^{*}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\bigcup_{(i,k)\in X_{0}^{j}\times X_{0}^{j}\,,i\neq j}M_{ik}^{j}\cup\bigcup_{(i,k)\in X_{1}^{j}\times X_{1}^{j}\,,i\neq j}M_{ik}^{j}\cup\bigcup_{(i,k)\in X_{0}^{j}\times X_{1}^{j}}M_{ik}^{j}\,, (5.32)

where Mikj=TUijUkjM_{ik}^{j}=T\cap\partial^{*}U_{i}^{j}\cap\partial^{*}U_{k}^{j} (an analogous decomposition of KK^{*} holds as well, and will be used in the following, but is not explicitly written for the sake of brevity). We now prove that

MikjEj(0),i,kX0j,ik,\displaystyle M_{ik}^{j}\subset E_{j}^{\scriptscriptstyle{(0)}}\,,\qquad\forall i,k\in X^{j}_{0}\,,i\neq k\,, (5.35)
MikjeEj,iX0j,kX1j,\displaystyle M_{ik}^{j}\subset\partial^{e}E_{j}\,,\qquad\forall i\in X^{j}_{0}\,,k\in X^{j}_{1}\,,
MikjEj(1),i,kX1j,ik.\displaystyle M_{ik}^{j}\subset E_{j}^{{\scriptscriptstyle{(1)}}}\,,\qquad\forall i,k\in X^{j}_{1}\,,i\neq k\,.

To prove (5.35) and (5.35): if iki\neq k, i,kX0ji,k\in X^{j}_{0}, and xMikjx\in M_{ik}^{j}, then (by |UijUkj|=0|U_{i}^{j}\cap U_{k}^{j}|=0) UijU_{i}^{j} and UkjU_{k}^{j} blow-up two complementary half-spaces at xx, an information that combined with the n+1\mathcal{L}^{n+1}-inclusion of UijUkjU_{i}^{j}\cup U_{k}^{j} in n+1Ej\mathbb{R}^{n+1}\setminus E_{j} implies

|Br(x)|+o(rn+1)=|Br(x)Uij|+|Br(x)Ukj||Br(x)Ej|,\displaystyle|B_{r}(x)|+{\rm o}(r^{n+1})=|B_{r}(x)\cap U_{i}^{j}|+|B_{r}(x)\cap U_{k}^{j}|\leq|B_{r}(x)\setminus E_{j}|\,,

that is, xEj(0)x\in E_{j}^{\scriptscriptstyle{(0)}}, thus proving (5.35); the proof of (5.35) is analogous.

To prove (5.35): if iX0ji\in X^{j}_{0}, kX1jk\in X^{j}_{1}, and xMikjx\in M_{ik}^{j}, then

|Br(x)Ej||Br(x)Ukj|=|Br(x)|2+o(rn+1),\displaystyle|B_{r}(x)\cap E_{j}|\geq|B_{r}(x)\cap U_{k}^{j}|=\frac{|B_{r}(x)|}{2}+{\rm o}(r^{n+1})\,,
|Br(x)Ej||Br(x)Uij|=|Br(x)|2+o(rn+1),\displaystyle|B_{r}(x)\setminus E_{j}|\geq|B_{r}(x)\cap U_{i}^{j}|=\frac{|B_{r}(x)|}{2}+{\rm o}(r^{n+1})\,,

so that xEj(0)x\not\in E_{j}^{\scriptscriptstyle{(0)}} and xEj(1)x\not\in E_{j}^{{\scriptscriptstyle{(1)}}}, i.e. xeEjx\in\partial^{e}E_{j}, that is (5.35).

With (5.35)–(5.35) at hand, we now prove that

TEj=n(i,k)X0j×X1jMikj,\displaystyle T\cap\partial^{*}E_{j}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\bigcup_{(i,k)\in X_{0}^{j}\times X_{1}^{j}}M_{ik}^{j}\,, (5.36)
KjEj(0)=n(i,k)X0j×X0j,kiMikj.\displaystyle K_{j}^{*}\cap E_{j}^{\scriptscriptstyle{(0)}}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\bigcup_{(i,k)\in X_{0}^{j}\times X_{0}^{j}\,,k\neq i}M_{ik}^{j}\,. (5.37)

(Analogous relations hold with KK^{*} and EE in place of KjK^{*}_{j} and EjE_{j}.)

To prove (5.36): By EjeEj\partial^{*}E_{j}\subset\partial^{e}E_{j} and (4.4) we find Ej(Uij)(1)=\partial^{*}E_{j}\cap(U_{i}^{j})^{{\scriptscriptstyle{(1)}}}=\varnothing for every i,ji,j; hence, since {(Uij)(1)}i{Uij}i\{(U_{i}^{j})^{{\scriptscriptstyle{(1)}}}\}_{i}\cup\{\partial^{*}U_{i}^{j}\}_{i} is an n\mathcal{H}^{n}-partition of TT, and by repeatedly applying (5.35), (5.35) and (5.35), we find

(i,k)X0j×X1jMikj\displaystyle\bigcup_{(i,k)\in X_{0}^{j}\times X_{1}^{j}}M_{ik}^{j} n\displaystyle\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset} TEj=ni(TEjUij)=ni,kMikjEj\displaystyle T\cap\partial^{*}E_{j}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\bigcup_{i}(T\cap\partial^{*}E_{j}\cap\partial^{*}U_{i}^{j})\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\bigcup_{i,k}M_{ik}^{j}\cap\partial^{*}E_{j}
=n\displaystyle\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=} (i,k)X0j×X1jMikjEj,\displaystyle\bigcup_{(i,k)\in X_{0}^{j}\times X_{1}^{j}}M_{ik}^{j}\cap\partial^{*}E_{j}\,,

which gives (5.36).

To prove (5.37): By (5.35), (5.35), and (5.35), MikjM_{ik}^{j} has empty intersection with Ej(0)E_{j}^{\scriptscriptstyle{(0)}} unless i,kX0ji,k\in X_{0}^{j}, in which case MikjM_{ik}^{j} is n\mathcal{H}^{n}-contained in Ej(0)E_{j}^{\scriptscriptstyle{(0)}}: hence,

(i,k)X0j×X0j,kiMikjnKjEj(0)=(i,k)X0j×X0j,kiEj(0)Mikj,\bigcup_{(i,k)\in X_{0}^{j}\times X_{0}^{j}\,,k\neq i}M_{ik}^{j}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}K_{j}^{*}\cap E_{j}^{\scriptscriptstyle{(0)}}=\bigcup_{(i,k)\in X_{0}^{j}\times X_{0}^{j}\,,k\neq i}E_{j}^{\scriptscriptstyle{(0)}}\cap M_{ik}^{j}\,,

that is (5.37).

With (5.36) and (5.37) at hand, we now prove the following perimeter formulas: for every open set ATA\subset T and every jj,

iX0jP(Uij;A)=n(AEj)+2n(AKjEj(0)),\displaystyle\sum_{i\in X^{j}_{0}}P(U_{i}^{j};A)=\mathcal{H}^{n}\big{(}A\cap\partial^{*}E_{j}\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap K^{*}_{j}\cap E_{j}^{\scriptscriptstyle{(0)}}\big{)}\,, (5.38)
iX1jP(Uij;A)=n(AEj)+2n(AKjEj(1)).\displaystyle\sum_{i\in X^{j}_{1}}P(U_{i}^{j};A)=\mathcal{H}^{n}\big{(}A\cap\partial^{*}E_{j}\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap K^{*}_{j}\cap E_{j}^{\scriptscriptstyle{(1)}}\big{)}\,. (5.39)

Analogously, for α=0,1\alpha=0,1,

iXαP(Ui;A)=n(AE)+2n(AKE(α)).\sum_{i\in X_{\alpha}}P(U_{i};A)=\mathcal{H}^{n}\big{(}A\cap\partial^{*}E\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap K^{*}\cap E^{\scriptscriptstyle{(\alpha)}}\big{)}\,. (5.40)

To prove (5.38) and (5.39): Indeed, by (5.36) and (5.37),

iX0jP(Uij;A)\displaystyle\sum_{i\in X_{0}^{j}}P(U_{i}^{j};A) =\displaystyle= (i,k)X0j×X1jn(AMikj)+iX0jkX0j{i}n(AMikj)\displaystyle\sum_{(i,k)\in X_{0}^{j}\times X_{1}^{j}}\mathcal{H}^{n}(A\cap M_{ik}^{j})+\sum_{i\in X_{0}^{j}}\sum_{k\in X_{0}^{j}\setminus\{i\}}\mathcal{H}^{n}(A\cap M_{ik}^{j})
=\displaystyle= n((i,k)X0j×X1jAMikj)+2n((i,k)X0j×X0j,ikAMikj)\displaystyle\mathcal{H}^{n}\Big{(}\bigcup_{(i,k)\in X_{0}^{j}\times X_{1}^{j}}A\cap M_{ik}^{j}\Big{)}+2\,\mathcal{H}^{n}\Big{(}\bigcup_{(i,k)\in X_{0}^{j}\times X_{0}^{j}\,,i\neq k}A\cap M_{ik}^{j}\Big{)}
=\displaystyle= n(AE)+2n(AKjEj(0)),\displaystyle\mathcal{H}^{n}(A\cap\partial^{*}E)+2\,\mathcal{H}^{n}\big{(}A\cap K_{j}^{*}\cap E_{j}^{\scriptscriptstyle{(0)}}\big{)}\,,

that is (5.38). The proof of (5.39) is analogous (since (5.39) is (5.38) applied to the complements of the EjE_{j}’s – recall indeed that ΩEj=Ω(ΩEj)\Omega\cap\partial^{*}E_{j}=\Omega\cap\partial^{*}(\Omega\setminus E_{j})).

Conclusion of the proof of (5.22) in the claim: We want to prove that K(T[s0]E(1))K^{*}\setminus(T[s_{0}]\cup E^{\scriptscriptstyle{(1)}}) is n\mathcal{H}^{n}-contained in KbkK_{\rm bk}. Since {E(0),E(1),E}\{E^{\scriptscriptstyle{(0)}},E^{\scriptscriptstyle{(1)}},\partial^{*}E\} is an n\mathcal{H}^{n}-partition of Ω\Omega, and ΩE\Omega\cap\partial^{*}E is contained in KbkK_{\rm bk}, looking back at the definition (5.4) of KbkK_{\rm bk} it is enough to show that

θn(μbk)(x)2 for n-a.e. x(KE(0))T[s0].\mbox{$\theta^{n}_{*}(\mu_{\rm bk})(x)\geq 2$ for $\mathcal{H}^{n}$-a.e. $x\in(K^{*}\cap E^{\scriptscriptstyle{(0)}})\setminus T[s_{0}]$}\,. (5.41)

To this end, we begin noticing that, if Y0Y_{0} is an arbitrary finite subset of X0X_{0}, then there is j(Y0)j(Y_{0}) such that Y0X0jY_{0}\subset X_{0}^{j} for every jj(Y0)j\geq j(Y_{0}); correspondingly,

iY0P(Ui;A)lim infjiY0P(Uij;A)lim infjiX0jP(Uij;A).\sum_{i\in Y_{0}}P(U_{i};A)\leq\liminf_{j\to\infty}\sum_{i\in Y_{0}}P(U_{i}^{j};A)\leq\liminf_{j\to\infty}\sum_{i\in X_{0}^{j}}P(U_{i}^{j};A)\,.

By arbitrariness of Y0Y_{0}, (5.40) with α=0\alpha=0, (5.38), and (4.11) (notice that the n\mathcal{H}^{n}-containment of the n\mathcal{H}^{n}-rectifiable set KjK^{*}_{j} into KjT[s0]K_{j}\cup T[s_{0}] is equivalent to its n\mathcal{H}^{n}-containment in (KjT[sj])=(Kj)T[sj]\mathcal{R}(K_{j}\cup T[s_{j}])=\mathcal{R}(K_{j})\cup T[s_{j}]) we conclude that, if ATA\subset T is open and such that cl(A)T[s0]=\mathrm{cl}\,(A)\cap T[s_{0}]=\varnothing, so that AT[sj]=A\cap T[s_{j}]=\varnothing for jj large enough, then

n(AE)+2n(AKE(0))\displaystyle\mathcal{H}^{n}\big{(}A\cap\partial^{*}E\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap K^{*}\cap E^{\scriptscriptstyle{(0)}}\big{)}
=iX0P(Ui;A)lim infjiX0jP(Uij;A)\displaystyle=\sum_{i\in X_{0}}P(U_{i};A)\leq\liminf_{j\to\infty}\sum_{i\in X_{0}^{j}}P(U_{i}^{j};A)
=lim infjn(AEj)+2n(AKjEj(0))\displaystyle=\liminf_{j\to\infty}\mathcal{H}^{n}\big{(}A\cap\partial^{*}E_{j}\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap K^{*}_{j}\cap E_{j}^{\scriptscriptstyle{(0)}}\big{)}
lim infjn(AEj)+2n(A((Kj)T[sj])Ej(0))\displaystyle\leq\liminf_{j\to\infty}\mathcal{H}^{n}\big{(}A\cap\partial^{*}E_{j}\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap\big{(}\mathcal{R}(K_{j})\cup T[s_{j}]\big{)}\cap E_{j}^{\scriptscriptstyle{(0)}}\big{)}
=lim infjn(AEj)+2n(A(Kj)Ej(0))μbk(cl(A)),\displaystyle=\liminf_{j\to\infty}\mathcal{H}^{n}\big{(}A\cap\partial^{*}E_{j}\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(0)}}\big{)}\leq\mu_{\rm bk}(\mathrm{cl}\,(A))\,, (5.42)

where we have used the definition (5.2) of μbk\mu_{\rm bk}. Now, if x(KE(0))T[s0]x\in(K^{*}\cap E^{\scriptscriptstyle{(0)}})\setminus T[s_{0}], then we we can apply (5.42) with A=Bs(x)A=B_{s}(x) and s>0s>0 such that cl(Bs(x))T[s0]=\mathrm{cl}\,(B_{s}(x))\cap T[s_{0}]=\varnothing, together with the fact that xE(0)x\in E^{\scriptscriptstyle{(0)}} implies n(Bs(x)E)=o(sn)\mathcal{H}^{n}(B_{s}(x)\cap\partial^{*}E)={\rm o}(s^{n}) as s0+s\to 0^{+}, to conclude that

μbk(cl(Bs(x)))\displaystyle\mu_{\rm bk}(\mathrm{cl}\,(B_{s}(x))) 2n(Bs(x)KE(0))+o(sn),as s0+.\displaystyle\geq 2\,\mathcal{H}^{n}\big{(}B_{s}(x)\cap K^{*}\cap E^{\scriptscriptstyle{(0)}}\big{)}+{\rm o}(s^{n})\,,\qquad\mbox{as $s\to 0^{+}$}\,. (5.43)

Since KE(0)K^{*}\cap E^{\scriptscriptstyle{(0)}} is an n\mathcal{H}^{n}-rectifiable set, and thus n(Bs(x)KE(0))=ωnsn+o(sn)\mathcal{H}^{n}\big{(}B_{s}(x)\cap K^{*}\cap E^{\scriptscriptstyle{(0)}}\big{)}=\omega_{n}\,s^{n}+{\rm o}(s^{n}) for n\mathcal{H}^{n}-a.e. xKE(0)x\in K^{*}\cap E^{\scriptscriptstyle{(0)}}, we deduce (5.41) from (5.43).

Conclusion of the proof of (5.23) in the claim: We want to prove the n\mathcal{H}^{n}-containment of KT[s0]K^{*}\setminus T[s_{0}] in KbdK_{\rm bd}. As in the proof of (5.22), combining Federer’s theorem (1.37) with the definition (5.5) of KbdK_{\rm bd}, we are left to prove that

θn(μbd)(x)2 for n-a.e. xK(T[s0]E).\mbox{$\theta_{*}^{n}(\mu_{\rm bd})(x)\geq 2$ for $\mathcal{H}^{n}$-a.e. $x\in K^{*}\setminus(T[s_{0}]\cup\partial^{*}E)$}\,. (5.44)

As proved in (5.42), if ATA\subset T is open and such that cl(A)T[s0]=\mathrm{cl}\,(A)\cap T[s_{0}]=\varnothing, then by exploiting (5.38) and (5.40) with α=0\alpha=0 we have

n(AE)+2n(AKE(0))\displaystyle\mathcal{H}^{n}\big{(}A\cap\partial^{*}E\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap K^{*}\cap E^{\scriptscriptstyle{(0)}}\big{)} (5.45)
lim infjn(AEj)+2n(A(Kj)Ej(0));\displaystyle\leq\liminf_{j\to\infty}\mathcal{H}^{n}\big{(}A\cap\partial^{*}E_{j}\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(0)}}\big{)}\,;

the same argument, this time based on (5.39) and (5.40) with α=1\alpha=1, also gives

n(AE)+2n(AKE(1))\displaystyle\mathcal{H}^{n}\big{(}A\cap\partial^{*}E\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap K^{*}\cap E^{\scriptscriptstyle{(1)}}\big{)} (5.46)
lim infjn(AEj)+2n(A(Kj)Ej(1));\displaystyle\leq\liminf_{j\to\infty}\mathcal{H}^{n}\big{(}A\cap\partial^{*}E_{j}\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(1)}}\big{)}\,;

and, finally, since ΩE\Omega\setminus\partial^{*}E is n\mathcal{H}^{n}-equivalent to Ω(E(0)E(1))\Omega\cap(E^{\scriptscriptstyle{(0)}}\cup E^{\scriptscriptstyle{(1)}}), the combination of (5.45) and (5.46) gives

n(AE)+2n(AKE)\displaystyle\mathcal{H}^{n}\big{(}A\cap\partial^{*}E\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap K^{*}\setminus\partial^{*}E\big{)} (5.47)
lim infjn(AEj)+2n(A(Kj)Ej)μbd(cl(A)),\displaystyle\leq\liminf_{j\to\infty}\mathcal{H}^{n}\big{(}A\cap\partial^{*}E_{j}\big{)}+2\,\mathcal{H}^{n}\big{(}A\cap\mathcal{R}(K_{j})\setminus\partial^{*}E_{j}\big{)}\leq\mu_{\rm bd}(\mathrm{cl}\,(A))\,,

where we have used the definition (5.3) of μbd\mu_{\rm bd}. Now, for n\mathcal{H}^{n}-a.e. xK(T[s0]E)x\in K^{*}\setminus(T[s_{0}]\cup\partial^{*}E) we have n(Br(x)E)=o(rn)\mathcal{H}^{n}(B_{r}(x)\cap\partial^{*}E)={\rm o}(r^{n}) and n(Br(x)KE)=ωnrn+o(rn)\mathcal{H}^{n}(B_{r}(x)\cap K^{*}\setminus\partial^{*}E)=\omega_{n}\,r^{n}+{\rm o}(r^{n}) as r0+r\to 0^{+}, as well as cl(Br(x))T[s0]=\mathrm{cl}\,(B_{r}(x))\cap T[s_{0}]=\varnothing for rr small enough, so that (5.47) with A=Br(x)A=B_{r}(x) readily implies (5.44). The proof of the claim, and thus of the theorem, is now complete. ∎

5.2. A second closure theorem

We now present a variant of the main arguments presented in this section and alternative closure theorem to Theorem 1.4. As already noticed, this second closure theorem, Theorem 5.1 below, will play a role only in the companion paper [MNR23a], where Plateau’s laws will be studied in the relation to the Allen–Cahn equation, so that this section can be omitted on a first reading focused on Gauss’ capillarity theory alone.

To introduce Theorem 1.4, let us consider the following question: given an n\mathcal{H}^{n}-finite set SS which is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, what parts of SS are essential to its 𝒞\mathcal{C}-spanning property? We already know from Lemma 2.2 that the unrectifiable part of SS is not necessary, since (S)\mathcal{R}(S) is also 𝒞\mathcal{C}-spanning. However, some parts of (S)\mathcal{R}(S) could be discarded too – indeed rectifiable sets can be “porous at every scale”, and thus completely useless from the point of view of achieving 𝒞\mathcal{C}-spanning. To make an example, consider the rectifiable set P2P\subset\mathbb{R}^{2} obtained by removing from [0,1][0,1] all the intervals (qiεi,qi+εi)(q_{i}-\varepsilon_{i},q_{i}+\varepsilon_{i}) where {qi}i\{q_{i}\}_{i} are the rational numbers in [0,1][0,1] and 2iεi=ε2\,\sum_{i}\varepsilon_{i}=\varepsilon for some given ε(0,1)\varepsilon\in(0,1): it is easily seen that PP is a rectifiable set with positive 1\mathcal{H}^{1}-measure in 2\mathbb{R}^{2}, contained in ×{0}\mathbb{R}\times\{0\}, which fails to essentially disconnect any stripe of the form (a,b)×(a,b)\times\mathbb{R} with (a,b)(0,1)(a,b)\subset\!\subset(0,1). Intuitively, if a set like PP stands as an isolated portion of SS, then (S)P\mathcal{R}(S)\setminus P should still be 𝒞\mathcal{C}-spanning.

We can formalize this idea as follows. Denoting as usual Ω=n+1𝐖\Omega=\mathbb{R}^{n+1}\setminus\mathbf{W}, we consider the open covering {Ωk}k\{\Omega_{k}\}_{k} of Ω\Omega defined by

{Ωk}k={Brmh(xm)}m,h,\{\Omega_{k}\}_{k}=\{B_{r_{mh}}(x_{m})\}_{m,h}\,, (5.48)

where {xm}m=n+1Ω\{x_{m}\}_{m}=\mathbb{Q}^{n+1}\cap\Omega and {rmh}h=(0,dist(xm,Ω))\{r_{mh}\}_{h}=\mathbb{Q}\cap(0,{\rm dist}(x_{m},\partial\Omega)). For every n\mathcal{H}^{n}-finite set SS we define the essential spanning part of SS in Ω\Omega as the Borel set

ESP(S)=kUBEP(S;Ωk)=k{ΩkiUi[Ωk]},{\rm ESP}(S)=\bigcup_{k}\,{\rm UBEP}(S;\Omega_{k})=\bigcup_{k}\,\Big{\{}\Omega_{k}\cap\bigcup_{i}\partial^{*}U_{i}[\Omega_{k}]\Big{\}}\,,

if {Ui[Ωk]}i\{U_{i}[\Omega_{k}]\}_{i} denotes the essential partition of Ωk\Omega_{k} induced by SS. Since each UBEP(S;Ωk){\rm UBEP}(S;\Omega_{k}) is a countable union of reduced boundaries and is n\mathcal{H}^{n}-contained in the n\mathcal{H}^{n}-finite set SS, we see that ESP(S){\rm ESP}(S) is always n\mathcal{H}^{n}-rectifiable. The idea is that by following the unions of boundaries of essential partitions induced by SS over smaller and smaller balls we are capturing all the parts of SS that may potentially contribute to achieve a spanning condition with respect to 𝐖\mathbf{W}. Thinking about Figure 1.5: the tendrils of SS appearing in panel (a) and not captured by UBEP(S;U){\rm UBEP}(S;U), will eventually be included into ESP(S){\rm ESP}(S) by considering UBEP{\rm UBEP}’s of SS relative to suitable subsets of UU. Another way to visualize the construction of ESP(S){\rm ESP}(S) is noticing that if Br(x)Bs(x)ΩB_{r}(x)\subset B_{s}(x)\subset\Omega, then

Br(x)UBEP(S;Bs(x))UBEP(S;Br(x)),B_{r}(x)\cap{\rm UBEP}(S;B_{s}(x))\subset{\rm UBEP}(S;B_{r}(x))\,,

which points to the monotonicity property behind the construction of ESP(S){\rm ESP}(S). Intuitively, we expect that

if SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, then ESP(S){\rm ESP}(S) is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} (5.49)

(where 𝒞\mathcal{C} is an arbitrary spanning class for 𝐖\mathbf{W}). This fact will proved in a moment as a particular case of Theorem 5.1 below.

Next, we introduce the notion of convergence behind our second closure theorem. Consider a sequence {Sj}j\{S_{j}\}_{j} of Borel subsets of Ω\Omega such that supjn(Sj)<\sup_{j}\mathcal{H}^{n}(S_{j})<\infty. If we denote by {Uij[Ωk]}i\{U_{i}^{j}[\Omega_{k}]\}_{i} the essential partition induced on Ωk\Omega_{k} by SjS_{j}, then a diagonal argument based on Lemma 2.3 shows the existence of a (not relabeled) subsequence in jj, and, for each kk, of a Borel partition {Ui[Ωk]}i\{U_{i}[\Omega_{k}]\}_{i} of Ωk\Omega_{k} such that {Uij[Ωk]}i\{U_{i}^{j}[\Omega_{k}]\}_{i} converges to {Ui[Ωk]}i\{U_{i}[\Omega_{k}]\}_{i} as jj\to\infty in the sense specified by (2.8). Since UBEP(Sj;Ωk)=ΩkiUij[Ωk]{\rm UBEP}(S_{j};\Omega_{k})=\Omega_{k}\cap\bigcup_{i}\partial^{*}U_{i}^{j}[\Omega_{k}], we call any set SS of the form111111The limit partition {Ui[Ωk]}i\{U_{i}[\Omega_{k}]\}_{i} appearing in (5.50) may not be the essential partition induced by SS on Ωk\Omega_{k} since the individual Ui[Ωk]U_{i}[\Omega_{k}], arising as L1L^{1}-limits, may fail to be essentially connected. This said, {Ui[Ωk]}i\{U_{i}[\Omega_{k}]\}_{i} is automatically a partition of Ωk\Omega_{k} induced by S0S_{0}.

S=k{ΩkiUi[Ωk]},S=\bigcup_{k}\,\Big{\{}\Omega_{k}\cap\bigcup_{i}\partial^{*}U_{i}[\Omega_{k}]\Big{\}}\,, (5.50)

a subsequential partition limit of {Sj}j\{S_{j}\}_{j} in Ω\Omega. Having in mind (5.49), it is natural to ask if the following property holds:

if Sj is 𝒞-spanning 𝐖 for each j,\displaystyle\mbox{if $S_{j}$ is $\mathcal{C}$-spanning $\mathbf{W}$ for each $j$}\,,
and S is a subsequential partition limit of {Sj}j in Ω,\displaystyle\mbox{and $S$ is a subsequential partition limit of $\{S_{j}\}_{j}$ in $\Omega$}\,,
then S is 𝒞-spanning 𝐖.\displaystyle\mbox{then $S$ is $\mathcal{C}$-spanning $\mathbf{W}$}\,. (5.51)

Our next theorem implies both (5.49) and (5.51) as particular cases (corresponding to be taking Ej=E_{j}=\varnothing and, respectively, Kj=SK_{j}=S and Kj=SjK_{j}=S_{j} for every jj).

Theorem 5.1 (Closure theorem for subsequential partition limits).

Let 𝐖\mathbf{W} be a closed set in n+1\mathbb{R}^{n+1}, 𝒞\mathcal{C} a spanning class for 𝐖\mathbf{W}, and {(Kj,Ej)}j\{(K_{j},E_{j})\}_{j} a sequence in 𝒦B\mathcal{K}_{\rm B} such that supjn(Kj)<\sup_{j}\mathcal{H}^{n}(K_{j})<\infty and KjEj(1)K_{j}\cup E_{j}^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} for every jj.

If S0S_{0} and E0E_{0} are, respectively, a subsequential partition limit of {Kj}j\{K_{j}\}_{j} in Ω\Omega and an L1L^{1}-subsequential limit of {Ej}j\{E_{j}\}_{j} (corresponding to a same not relabeled subsequence in jj), and we set

K0=(ΩE0)S0,K_{0}=(\Omega\cap\partial^{*}E_{0})\cup S_{0}\,,

then (K0,E0)𝒦B(K_{0},E_{0})\in\mathcal{K}_{\rm B} and K0E0(1)K_{0}\cup E_{0}^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}.

Proof.

Since ΩE0K0\Omega\cap\partial^{*}E_{0}\subset K_{0} by definition of K0K_{0} we trivially have (K0,E0)𝒦B(K_{0},E_{0})\in\mathcal{K}_{\rm B}. Aiming to prove that K0E0(1)K_{0}\cup E_{0}^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, we fix (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}), and define s0s_{0}, sjs_{j}, {Uij}i\{U_{i}^{j}\}_{i} and {Ui}i\{U_{i}\}_{i} exactly as in part I of the proof of Theorem 1.4. Thanks to Theorem 4.1 and by arguing as in part II of the proof of Theorem 1.4, we have reduced to prove that

K(T[s0]E(1)) is n-contained in K0.\mbox{$K^{*}\setminus(T[s_{0}]\cup E^{\scriptscriptstyle{(1)}})$ is $\mathcal{H}^{n}$-contained in $K_{0}$}\,. (5.52)

By Federer’s theorem (1.37) and since ΩEK0\Omega\cap\partial^{*}E\subset K_{0} it is enough to prove

(KE(0))T[s0] is n-contained in S0,\mbox{$(K^{*}\cap E^{\scriptscriptstyle{(0)}})\setminus T[s_{0}]$ is $\mathcal{H}^{n}$-contained in $S_{0}$}\,,

and, thanks to the construction of S0S_{0}, we shall actually be able to prove

KT[s0] is n-contained in S0.\mbox{$K^{*}\setminus T[s_{0}]$ is $\mathcal{H}^{n}$-contained in $S_{0}$}\,. (5.53)

To this end let us pick kk such that ΩkT\Omega_{k}\subset\!\subset T and ΩkT[s0]=\Omega_{k}\cap T[s_{0}]=\emptyset. Then, for jj(k)j\geq j(k), we have ΩkT[sj]=\Omega_{k}\cap T[s_{j}]=\varnothing, so that

ΩkUBEP(KjT[sj];T)UBEP(KjT[sj];Ωk)=UBEP(Kj;Ωk).\Omega_{k}\cap{\rm UBEP}\big{(}K_{j}\cup T[s_{j}];T\big{)}\subset{\rm UBEP}\big{(}K_{j}\cup T[s_{j}];\Omega_{k}\big{)}={\rm UBEP}\big{(}K_{j};\Omega_{k}\big{)}\,.

Since {Uij}i\{U_{i}^{j}\}_{i} is the essential partition of TT induced by KjT[sj]K_{j}\cup T[s_{j}], if {Umj[Ωk]}m\{U_{m}^{j}[\Omega_{k}]\}_{m} is the essential partition of Ωk\Omega_{k} induced by KjK_{j}, we have just claimed that, for every ii and jj(k)j\geq j(k),

ΩkUijΩkmUmj[Ωk].\Omega_{k}\cap\partial^{*}U_{i}^{j}\subset\Omega_{k}\cap\bigcup_{m}\partial^{*}U_{m}^{j}[\Omega_{k}]\,. (5.54)

Since {Umj[Ωk]}m\{U_{m}^{j}[\Omega_{k}]\}_{m} is a Lebesgue partition of Ωk\Omega_{k} into essentially connected sets, by (5.54) the indecomposable components of ΩkUij\Omega_{k}\cap U_{i}^{j} must belong to {Umj[Ωk]}m\{U_{m}^{j}[\Omega_{k}]\}_{m}. In other words, for each ii and each jj(k)j\geq j(k) there is M(k,i,j)M(k,i,j) such that

ΩkUij=mM(k,i,j)Umj[Ωk].\Omega_{k}\cap U_{i}^{j}=\bigcup_{m\in M(k,i,j)}U_{m}^{j}[\Omega_{k}]\,.

As a consequence of UijUiU_{i}^{j}\to U_{i} and of Umj[Ωk]Um[Ωk]U_{m}^{j}[\Omega_{k}]\to U_{m}[\Omega_{k}] as jj\to\infty we find that, for a set of indexes M(k,i)M(k,i), it must be

ΩkUi=mM(k,i)Um[Ωk],\Omega_{k}\cap U_{i}=\bigcup_{m\in M(k,i)}U_{m}[\Omega_{k}]\,,

and therefore

ΩkUinmM(k,i)Um[Ωk]S0.\Omega_{k}\cap\partial^{*}U_{i}\,\,\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}\,\,\bigcup_{m\in M(k,i)}\partial^{*}U_{m}[\Omega_{k}]\subset S_{0}\,.

Since we have proved this inclusion for every ii and for every kk such that ΩkT\Omega_{k}\subset\!\subset T with ΩkT[s0]=\Omega_{k}\cap T[s_{0}]=\emptyset, it follows that KT[s0]K^{*}\setminus T[s_{0}] is n\mathcal{H}^{n}-contained in S0S_{0}, that is (5.53). ∎

6. Existence of minimizers and convergence to Plateau’s problem (Theorem 1.5)

In this section we prove two main results: the first one (Theorem 6.1) concerns the equivalence of Harrison–Pugh Plateau’s problem \ell with its measure theoretic reformulation B\ell_{\rm B} (see (1.21)); the second (Theorem 1.5) is a very refined version of Theorem 1.5.

Theorem 6.1 (Existence for B\ell_{\rm B} and =B\ell=\ell_{\rm B}).

If 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} is closed, 𝒞\mathcal{C} is a spanning class for 𝐖\mathbf{W}, and the Harrison–Pugh formulation of the Plateau problem

=inf{n(S):S is a closed subset ΩS is 𝒞-spanning 𝐖}\displaystyle\ell=\inf\big{\{}\mathcal{H}^{n}(S):\mbox{$S$ is a closed subset $\Omega$, $S$ is $\mathcal{C}$-spanning $\mathbf{W}$}\big{\}}

is finite, then the problem

B=inf{n(S):S is a Borel subset ΩS is 𝒞-spanning 𝐖}\displaystyle\ell_{\rm B}=\inf\big{\{}\mathcal{H}^{n}(S):\mbox{$S$ is a Borel subset $\Omega$, $S$ is $\mathcal{C}$-spanning $\mathbf{W}$}\big{\}}

admits minimizers, and given any minimizer SS for B\ell_{\rm B}, there exists relatively closed SS^{*} which is n\mathcal{H}^{n}-equivalent to SS and a minimizer for \ell. In particular, =B\ell=\ell_{\rm B}.

Theorem 6.2 (Theorem 1.5 refined).

If 𝐖\mathbf{W} is a compact set in n+1\mathbb{R}^{n+1} and 𝒞\mathcal{C} is a spanning class for 𝐖\mathbf{W} such that <\ell<\infty, then for every v>0v>0 there exist minimizers (K,E)(K,E) of Ψbk(v)\Psi_{\rm bk}(v). Moreover,

(i): if (K,E)(K_{*},E_{*}) is a minimizer of Ψbk(v)\Psi_{\rm bk}(v), then there is (K,E)𝒦(K,E)\in\mathcal{K} such that KK is n\mathcal{H}^{n}-equivalent to KK^{*}, EE is Lebesgue equivalent to EE_{*}, (K,E)(K,E) is a minimizer of Ψbk(v)\Psi_{\rm bk}(v), both EE and KK are bounded, KEK\cup E is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, KE(1)=K\cap E^{\scriptscriptstyle{(1)}}=\varnothing, and there is λ\lambda\in\mathbb{R} such that

λΩEXνE𝑑n=ΩEdivKX𝑑n+2KE(0)divKX𝑑n,\displaystyle\lambda\int_{\Omega\cap\partial^{*}E}X\cdot\nu_{E}\,d\mathcal{H}^{n}=\int_{\Omega\cap\partial^{*}E}\mathrm{div}^{K}\,X\,d\mathcal{H}^{n}+2\int_{K\cap E^{\scriptscriptstyle{(0)}}}\mathrm{div}^{K}\,X\,d\mathcal{H}^{n}\,, (6.1)
XCc1(n+1;n+1)with XνΩ=0 on Ω,\displaystyle\hskip 65.44142pt\forall X\in C^{1}_{c}(\mathbb{R}^{n+1};\mathbb{R}^{n+1})\quad\mbox{with $X\cdot\nu_{\Omega}=0$ on $\partial\Omega$}\,,

and there are positive constants c=c(n)c=c(n) and r1=r1(K,E)r_{1}=r_{1}(K,E) such that

|EBρ(y)|(1c)ωn+1ρn+1,|E\cap B_{\rho}(y)|\leq(1-c)\,\omega_{n+1}\,\rho^{n+1}\,, (6.2)

for every yΩEy\in\Omega\cap\partial E and ρ<min{r1,dist(y,𝐖)}\rho<\min\{r_{1},{\rm dist}(y,\mathbf{W})\}; under the further assumption that 𝐖\partial\mathbf{W} is C2C^{2}, then there is positive r0=r0(n,𝐖,|λ|)r_{0}=r_{0}(n,\mathbf{W},|\lambda|) such that

n(KBr(x))crn\mathcal{H}^{n}(K\cap B_{r}(x))\geq c\,r^{n} (6.3)

for every xcl(K)x\in\mathrm{cl}\,(K) and r<r0r<r_{0};

(ii): if (Kj,Ej)(K_{j},E_{j}) is a sequence of minimizers for Ψbk(vj)\Psi_{\rm bk}(v_{j}) with vj0+v_{j}\to 0^{+}, then there exists a minimizer SS of \ell such that, up to extracting subsequences, as Radon measures in Ω\Omega,

n (ΩEj)+2n (KjEj(0))2n S,\displaystyle\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E_{j})+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K_{j}\cap E_{j}^{{\scriptscriptstyle{(0)}}})\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}2\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S\,, (6.4)

as jj\to\infty. In particular, Ψbk(v)2=Ψbk(0)\Psi_{\rm bk}(v)\to 2\,\ell=\Psi_{\rm bk}(0) as v0+v\to 0^{+}.

Proof of Theorem 6.1.

By Theorem A.1, if <\ell<\infty, then B<\ell_{\rm B}<\infty. Let now {Sj}j\{S_{j}\}_{j} be a minimizing sequence for B\ell_{\rm B}, then {(Sj,)}j\{(S_{j},\varnothing)\}_{j} is a sequence in 𝒦B\mathcal{K}_{\rm B} satisfying (5.1). By Theorem 1.4, we find a Borel set SS which is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} and is such that

2lim infjn(Sj)=lim infjbk(Sj,)bk(S,)=2n(S).2\,\liminf_{j\to\infty}\mathcal{H}^{n}(S_{j})=\liminf_{j\to\infty}\mathcal{F}_{\rm bk}(S_{j},\varnothing)\geq\mathcal{F}_{\rm bk}(S,\varnothing)=2\,\mathcal{H}^{n}(S)\,.

This shows that SS is a minimizer of B\ell_{\rm B}. By Lemma 2.2, SS is n\mathcal{H}^{n}-rectifiable, for, otherwise, (S)\mathcal{R}(S) would be admissible for B\ell_{\rm B} and have strictly less area than SS. We conclude the proof by showing that up to modifications on a n\mathcal{H}^{n}-null set, SS is relatively closed in Ω\Omega (and thus is a minimizer of \ell too). Indeed the property of being 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} is preserved under diffeomorphism ff with {fid}Ω\{f\neq{\rm id}\,\}\subset\!\subset\Omega. In particular, n(S)n(f(S))\mathcal{H}^{n}(S)\leq\mathcal{H}^{n}(f(S)) for every such ff, so that the multiplicity one rectifiable varifold VS=𝐯𝐚𝐫(S,1)V_{S}=\mathbf{var}\,(S,1) associated to SS is stationary. By a standard application of the monotonicity formula, we can find SS^{*} n\mathcal{H}^{n}-equivalent to SS such that SS^{*} is relative closed in Ω\Omega. Since n(S)=n(S)\mathcal{H}^{n}(S)=\mathcal{H}^{n}(S^{*}) and 𝒞\mathcal{C}-spanning is preserved under n\mathcal{H}^{n}-null modifications, we conclude the proof. ∎

Proof of Theorem 6.2.

Step one: We prove conclusion (i). To this end, let (K,E)𝒦B(K_{*},E_{*})\in\mathcal{K}_{\rm B} be a minimizer of Ψbk(v)\Psi_{\rm bk}(v). Clearly, ((K),E)𝒦B(\mathcal{R}(K_{*}),E_{*})\in\mathcal{K}_{\rm B} is such that (K)E(1)\mathcal{R}(K_{*})\cup E^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} (thanks to Theorem 3.1/Remark 3.2) and bk((K),E)bk(K,E)\mathcal{F}_{\rm bk}(\mathcal{R}(K_{*}),E_{*})\leq\mathcal{F}_{\rm bk}(K_{*},E_{*}). In particular, ((K),E)(\mathcal{R}(K_{*}),E_{*}) is a minimizer of Ψbk(v)\Psi_{\rm bk}(v), and energy comparison between ((K),E)(\mathcal{R}(K_{*}),E_{*}) and ((K)E(1),E)(\mathcal{R}(K_{*})\setminus E_{*}^{\scriptscriptstyle{(1)}},E_{*}) (which is also a competitor for Ψbk(v)\Psi_{\rm bk}(v)) proves that

n((K)E(1))=0.\mathcal{H}^{n}(\mathcal{R}(K_{*})\cap E_{*}^{\scriptscriptstyle{(1)}})=0\,. (6.5)

Since “𝒞\mathcal{C}-spanning 𝐖\mathbf{W}” is preserved under diffeomorphisms, by a standard first variation argument (see, e.g. [KMS22a, Appendix C]) wee see that ((K),E)(\mathcal{R}(K_{*}),E_{*}) satisfies (6.1) for some λ\lambda\in\mathbb{R}. In particular, the integer nn-varifold V=var((K),θ)V={\rm var}(\mathcal{R}(K_{*}),\theta), with multiplicity function θ=2\theta=2 on (K)E(0)\mathcal{R}(K_{*})\cap E_{*}^{\scriptscriptstyle{(0)}} and θ=1\theta=1 on ΩE\Omega\cap\partial^{*}E_{*}, has bounded mean curvature in Ω\Omega, and thus satisfies V(Br(x))c(n)rn\|V\|(B_{r}(x))\geq c(n)\,r^{n} for every xKx\in K and r<min{r0,dist(x,𝐖)}r<\min\{r_{0},{\rm dist}(x,\mathbf{W})\}, where r0=r0(n,|λ|)r_{0}=r_{0}(n,|\lambda|) and, by definition,

K:=ΩsptV.K:=\Omega\cap{\rm spt}V\,.

In particular, since (6.5) implies V2n  (K)\|V\|\leq 2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}\mathcal{R}(K^{*}) , we conclude (e.g. by [Mag12, Corollary 6.4]) that KK is n\mathcal{H}^{n}-equivalent to (K)\mathcal{R}(K_{*}), and is thus n\mathcal{H}^{n}-rectifiable and relatively closed in Ω\Omega. Now let

E={xΩ:r<dist(x,𝐖) s.t. |EBr(x)|=|Br(x)|},E=\big{\{}x\in\Omega:\mbox{$\exists\,\,r<{\rm dist}(x,\mathbf{W})$ s.t. $|E_{*}\cap B_{r}(x)|=|B_{r}(x)|$}\big{\}}\,,

so that, trivially, EE is an open subset of Ω\Omega with EE(1)E\subset E_{*}^{\scriptscriptstyle{(1)}}. By applying (1.35) to EE_{*}, and by noticing that if xΩEx\in\Omega\setminus E then |EBr(x)|<|Br(x)||E_{*}\cap B_{r}(x)|<|B_{r}(x)| for every r>0r>0, and that if xΩcl(E)x\in\Omega\cap\mathrm{cl}\,(E) then |EBr(x)|>0|E_{*}\cap B_{r}(x)|>0 for every r>0r>0, we see that

ΩE{xΩ:0<|EBr(x)|<|Br(x)|r>0}=Ωcl(E).\Omega\cap\partial E\,\,\subset\,\,\big{\{}x\in\Omega:0<|E_{*}\cap B_{r}(x)|<|B_{r}(x)|\,\,\forall r>0\big{\}}\,\,=\,\,\Omega\cap\mathrm{cl}\,(\partial^{*}E_{*})\,. (6.6)

Since Vn  (ΩE)\|V\|\geq\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E_{*}) and n(Br(x)E)=ωnrn+o(rn)\mathcal{H}^{n}(B_{r}(x)\cap\partial^{*}E)=\omega_{n}\,r^{n}+{\rm o}(r^{n}) as r0+r\to 0^{+} for every xΩEx\in\Omega\cap\partial^{*}E, we see that ΩEΩsptV=K\Omega\cap\partial^{*}E_{*}\subset\Omega\cap{\rm spt}\|V\|=K, and since KK is relatively closed in Ω\Omega, we have Ωcl(E)K\Omega\cap\mathrm{cl}\,(\partial^{*}E_{*})\subset K, and so ΩEK\Omega\cap\partial E\subset K. In particular, EE is of finite perimeter, and thus by applying (1.35) to EE,

Ωcl(E)={xΩ:0<|EBr(x)|<|Br(x)|r>0}ΩE.\Omega\cap\mathrm{cl}\,(\partial^{*}E)\,\,=\,\,\big{\{}x\in\Omega:0<|E\cap B_{r}(x)|<|B_{r}(x)|\,\,\forall r>0\big{\}}\,\,\subset\,\,\Omega\cap\partial E\,. (6.7)

Finally, if there is x(ΩE(1))Ex\in(\Omega\cap E_{*}^{\scriptscriptstyle{(1)}})\setminus E, then it must be 0<|EBr(x)|<|Br(x)|0<|E_{*}\cap B_{r}(x)|<|B_{r}(x)| for every r>0r>0, and thus xΩcl(E)Kx\in\Omega\cap\mathrm{cl}\,(\partial^{*}E_{*})\subset K. However, we claim that for every xΩcl(E)x\in\Omega\cap\mathrm{cl}\,(\partial^{*}E_{*}) and r<min{r,dist(x,𝐖)}r<\min\{r_{*},{\rm dist}(x,\mathbf{W})\} (with r=r(K,E)r_{*}=r_{*}(K_{*},E_{*})) it holds

|Br(x)E|(1c)ωn+1rn+1,|B_{r}(x)\cap E_{*}|\leq(1-c)\,\omega_{n+1}\,r^{n+1}\,, (6.8)

in contradiction with xE(1)x\in E^{\scriptscriptstyle{(1)}}; this proves that ΩE(1)E\Omega\cap E_{*}^{\scriptscriptstyle{(1)}}\subset E, and thus that EE_{*} and EE are Lebesgue equivalent. Combining the latter information with (6.6) and (6.7) we conclude that Ωcl(E)=ΩEK\Omega\cap\mathrm{cl}\,(\partial^{*}E)=\Omega\cap\partial E\subset K and conclude the proof of (K,E)𝒦(K,E)\in\mathcal{K} – conditional to proving (6.8).

To prove (6.8), let us fix xΩcl(E)x\in\Omega\cap\mathrm{cl}\,(\partial^{*}E_{*}) and set u(r)=|Br(x)E|u(r)=|B_{r}(x)\setminus E_{*}|, so that, for a.e. r>0r>0 we have

u(r)=n(E(0)Br(x)),P(Br(x)E)=u(r)+P(E;Br(x)).\displaystyle u^{\prime}(r)=\mathcal{H}^{n}(E_{*}^{\scriptscriptstyle{(0)}}\cap\partial B_{r}(x))\,,\qquad P(B_{r}(x)\setminus E_{*})=u^{\prime}(r)+P(E_{*};B_{r}(x))\,. (6.9)

Since |E|=v>0|E_{*}|=v>0, we have n(ΩE)>0\mathcal{H}^{n}(\Omega\cap\partial^{*}E_{*})>0, therefore there must be y1,y2ΩEy_{1},y_{2}\in\Omega\cap\partial^{*}E_{*} with |y1y2|>4r|y_{1}-y_{2}|>4r_{*} for some rr_{*} depending on EE_{*}. In particular there is i{1,2}i\in\{1,2\} such that Br(x)Br(yi)=B_{r_{*}}(x)\cap B_{r_{*}}(y_{i})=\varnothing, and we set y=yiy=y_{i}. Since yiΩEy_{i}\in\Omega\cap\partial^{*}E_{*}, there is w>0w_{*}>0 and smooth maps Φ:Ω×(w,w)Ω\Phi:\Omega\times(-w_{*},w_{*})\to\Omega such that Φ(,w)\Phi(\cdot,w) is a diffeomorphism of Ω\Omega with {Φ(,w)Id}Br(y)\{\Phi(\cdot,w)\neq{\rm Id}\,\}\subset\!\subset B_{r_{*}}(y), and

|Φ(E,w)|=|E|w,P(Φ(E,w);Br(y))P(E,Br(y))(1+2|λ||w|),|\Phi(E_{*},w)|=|E_{*}|-w\,,\qquad P(\Phi(E_{*},w);B_{r_{*}}(y))\leq P(E_{*},B_{r_{*}}(y))(1+2\,|\lambda|\,|w|)\,, (6.10)

for every |w|<w|w|<w_{*}. We then consider r1r_{1} such that |Br1|<w|B_{r_{1}}|<w_{*}, so that for every r<min{r1,dist(x,𝐖)}r<\min\{r_{1},{\rm dist}(x,\mathbf{W})\} we have 0u(r)<w0\leq u(r)<w_{*}, and thus we can define

(Kr,Er)=(Φu(r)(KBr(x)),Φu(r)(EBr(x))).(K_{r},E_{r})=\Big{(}\Phi^{u(r)}\big{(}K\cup\partial B_{r}(x)\big{)},\Phi^{u(r)}\big{(}E_{*}\cup B_{r}(x)\big{)}\Big{)}\,.

Since Φu(r)\Phi^{u(r)} is a diffeomorphism, we have ΩErKr\Omega\cap\partial^{*}E_{r}\subset K_{r}, and by the first relation in (6.10) and Φu(r)=Id\Phi^{u(r)}={\rm Id}\, on ΩBr(y)\Omega\setminus B_{r_{*}}(y), we get

|Er||E|=|Br(x)||Br(x)E|+|Φu(r)(E)Br(y)||EBr(y)|=u(r)u(r)=0.|E_{r}|-|E|=|B_{r}(x)|-|B_{r}(x)\cap E_{*}|+|\Phi^{u(r)}(E_{*})\cap B_{r_{*}}(y)|-|E_{*}\cap B_{r_{*}}(y)|=u(r)-u(r)=0\,.

Hence bk(K,E)bk(Kr,Er)\mathcal{F}_{\rm bk}(K_{*},E_{*})\leq\mathcal{F}_{\rm bk}(K_{r},E_{r}), from which we deduce

P(E;Br(x))+P(E;Br(y))+2n(KE(0)Br(x))\displaystyle P(E;B_{r}(x))+P(E;B_{r_{*}}(y))+2\,\mathcal{H}^{n}(K_{*}\cap E_{*}^{\scriptscriptstyle{(0)}}\cap B_{r}(x))
n(Br(x)E(0))+P(Φu(r)(E);Br(y))u(r)+P(E,Br(y))(1+2|λ||w|);\displaystyle\leq\mathcal{H}^{n}(B_{r}(x)\cap E^{\scriptscriptstyle{(0)}})+P(\Phi^{u(r)}(E_{*});B_{r_{*}}(y))\leq u^{\prime}(r)+P(E_{*},B_{r_{*}}(y))(1+2\,|\lambda|\,|w|)\,;

where we have used (6.9) and (6.10); by adding up u(r)u^{\prime}(r) on both sides of the inequality, and using (6.9) again, we find that

c(n)u(r)n/(n+1)P(Br(x)E)2u(r)+2|λ|Ψbk(v)u(r),c(n)\,u(r)^{n/(n+1)}\leq P(B_{r}(x)\setminus E_{*})\leq 2\,u^{\prime}(r)+2\,|\lambda|\,\Psi_{\rm bk}(v)\,u(r)\,,

for a.e. r<min{r1,dist(x,𝐖)}r<\min\{r_{1},{\rm dist}(x,\mathbf{W})\}; since, by (6.6), xΩcl(E)x\in\Omega\cap\mathrm{cl}\,(\partial^{*}E_{*}) implies u(r)>0u(r)>0 for every r>0r>0, we can apply a standard ODE argument to conclude that (6.8) holds true.

We now prove the remaining assertions in statement (i). First of all, when 𝐖\partial\mathbf{W} is C2C^{2}, we can argue similarly to [KMS22b, Theorem 4.1] to deduce from the modified monotonicity formula of Kagaya and Tonegawa [KT17] that the area lower bound in (6.3) holds for every xcl(K)x\in\mathrm{cl}\,(K) and every r<r0r<r_{0}. The validity of the volume upper bound in (6.2) is immediate from (6.8) and the Lebesgue equivalence of EE_{*} and EE. The monotonicity formula for VV combined with n(ΩK)<\mathcal{H}^{n}(\Omega\cap K)<\infty implies of course that VV has bounded support. Having proved that KK is bounded, |E|<|E|<\infty and ΩEK\Omega\cap\partial E\subset K imply that EE is bounded too. Since (K)\mathcal{R}(K_{*}) and KK are n\mathcal{H}^{n}-equivalent, we have that KE(1)K\cup E_{*}^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}. It turns out that KE(1)K\cup E^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} too, since EE and EE_{*} are Lebesgue equivalent and of finite perimeter, therefore such that E(1)E^{\scriptscriptstyle{(1)}} and E(1)E_{*}^{\scriptscriptstyle{(1)}} are n\mathcal{H}^{n}-equivalent. In fact, on noticing that Ω(E(1)E)ΩEK\Omega\cap(E^{\scriptscriptstyle{(1)}}\setminus E)\subset\Omega\cap\partial E\subset K, we see that KE(1)=KEK\cup E^{\scriptscriptstyle{(1)}}=K\cup E, so that KEK\cup E is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, as claimed.

Finally, we prove that KE(1)=K\cap E^{\scriptscriptstyle{(1)}}=\varnothing. We first notice that, since EΩE\subset\Omega is open and K=ΩsptVK=\Omega\cap{\rm spt}\,V with V2n  (K)\|V\|\leq 2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}\mathcal{R}(K^{*}), if KEK\cap E\neq\emptyset, then n((K)E)>0\mathcal{H}^{n}(\mathcal{R}(K_{*})\cap E)>0; and since EE(1)E\subset E_{*}^{\scriptscriptstyle{(1)}} by construction, we arrive at a contradiction with (6.5). Hence, KE=K\cap E=\varnothing. Now, if xKE(1)x\in K\cap E^{\scriptscriptstyle{(1)}}, then, by (6.2), xΩEx\not\in\Omega\cap\partial E; combining this with KE=K\cap E=\varnothing, we find KE(1)Ωcl(E)E(0)K\cap E^{\scriptscriptstyle{(1)}}\subset\Omega\setminus\mathrm{cl}\,(E)\subset E^{\scriptscriptstyle{(0)}}, and thus KE(1)=K\cap E^{\scriptscriptstyle{(1)}}=\varnothing.

Step two: For every v10v_{1}\geq 0 and v2>0v_{2}>0 we have

Ψbk(v1+v2)Ψbk(v1)+(n+1)ωn+11/(n+1)v2n/(n+1).\Psi_{\rm bk}(v_{1}+v_{2})\leq\Psi_{\rm bk}(v_{1})+(n+1)\,\omega_{n+1}^{1/(n+1)}\,v_{2}^{n/(n+1)}\,. (6.11)

Since Ψbk(0)=2<\Psi_{\rm bk}(0)=2\,\ell<\infty, (6.11) implies in particular that Ψbk(v)<\Psi_{\rm bk}(v)<\infty for every v>0v>0 (just take v1=0v_{1}=0 and v2=vv_{2}=v).

Indeed, let (K1,E1)(K_{1},E_{1}) be a competitor in Ψbk(v1)\Psi_{\rm bk}(v_{1}) and let {Brj(xj)}j\{B_{r_{j}}(x_{j})\}_{j} be a sequence of balls with |xj||x_{j}|\to\infty and |E1Brj(xj)|=v1+v2|E_{1}\cup B_{r_{j}}(x_{j})|=v_{1}+v_{2} for every jj. Setting for the sake of brevity Bj=Brj(xj)B_{j}=B_{r_{j}}(x_{j}), sine (E1Bj)\partial^{*}(E_{1}\cup B_{j}) is n\mathcal{H}^{n}-contained in (E1)Bj(\partial^{*}E_{1})\cup\partial B_{j} we have that (K2,E2)(K_{2},E_{2}), with K2=K1BjK_{2}=K_{1}\cup\partial B_{j} and E2=E1BjE_{2}=E_{1}\cup B_{j}, is a competitor of Ψbk(v1+v2)\Psi_{\rm bk}(v_{1}+v_{2}). Since BjE2(0)=\partial B_{j}\cap E_{2}^{\scriptscriptstyle{(0)}}=\varnothing implies E2(0)E1(0)BjE_{2}^{\scriptscriptstyle{(0)}}\subset E_{1}^{\scriptscriptstyle{(0)}}\setminus\partial B_{j}, we find that

Ψbk(v1+v2)\displaystyle\Psi_{\rm bk}(v_{1}+v_{2}) \displaystyle\leq 2n(K2E2(0))+n(ΩE2)\displaystyle 2\,\mathcal{H}^{n}\big{(}K_{2}\cap E_{2}^{\scriptscriptstyle{(0)}})+\mathcal{H}^{n}(\Omega\cap\partial^{*}E_{2})
\displaystyle\leq 2n(K1E1(0)Bj)+n(ΩE1)+n(Bj)\displaystyle 2\,\mathcal{H}^{n}(K_{1}\cap E_{1}^{\scriptscriptstyle{(0)}}\setminus\partial B_{j})+\mathcal{H}^{n}(\Omega\cap\partial^{*}E_{1})+\mathcal{H}^{n}(\partial B_{j})
\displaystyle\leq bk(K1,E1)+(n+1)ωn+11/(n+1)|Bj|n/(n+1).\displaystyle\mathcal{F}_{\rm bk}(K_{1},E_{1})+(n+1)\,\omega_{n+1}^{1/(n+1)}\,|B_{j}|^{n/(n+1)}\,.

Since |xj||x_{j}|\to\infty, |E1|=v1|E_{1}|=v_{1}, and |E1Brj(xj)|=v1+v2|E_{1}\cup B_{r_{j}}(x_{j})|=v_{1}+v_{2} imply |Bj|v2|B_{j}|\to v_{2}, we conclude by arbitrariness of (K1,E1)(K_{1},E_{1}).

Step three: Now let {(Kj,Ej)}j\{(K_{j},E_{j})\}_{j} be a minimizing sequence for Ψbk(v)\Psi_{\rm bk}(v). Since Ψbk(v)<\Psi_{\rm bk}(v)<\infty, assumption (5.1) of Theorem 1.4 holds. Therefore there is (K,E)𝒦B(K,E)\in\mathcal{K}_{\rm B} with KE(1)K\cup E^{{\scriptscriptstyle{(1)}}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} and such that, up to extracting subsequences,

limj|(EjΔE)BR|=0R>0,lim infjbk(Kj,Ej)bk(K,E);\lim_{j\to\infty}|(E_{j}\Delta E)\cap B_{R}|=0\quad\forall R>0\,,\qquad\liminf_{j\to\infty}\mathcal{F}_{\rm bk}(K_{j},E_{j})\geq\mathcal{F}_{\rm bk}(K,E)\,; (6.12)

actually, to be more precise, if μ\mu denotes the weak-star limit of n  (ΩEj)+2n  ((Kj)Ej(0))\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E_{j})+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(0)}}) in Ω\Omega, then

μ2n  (KE(0))+n  (ΩE).\mu\geq 2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K\cap E^{\scriptscriptstyle{(0)}})+\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E)\,. (6.13)

We claim that

(K,E) is a minimizer of Ψbk(|E|).\mbox{$(K,E)$ is a minimizer of $\Psi_{\rm bk}(|E|)$}\,.

(Notice that, at this stage of the argument, we are not excluding that v:=v|E|v^{*}:=v-|E| is positive, nor that |E|=0|E|=0.) Taking into account (6.11), to prove the claim it suffices to show that

Ψbk(v)bk(K,E)+(n+1)ωn+11/(n+1)(v)n/(n+1).\displaystyle\Psi_{\rm bk}(v)\geq\mathcal{F}_{\rm bk}(K,E)+(n+1)\,\omega_{n+1}^{1/(n+1)}\,(v^{*})^{n/(n+1)}\,. (6.14)

To see this, we start noticing that, given any sequence {rj}j\{r_{j}\}_{j} with rjr_{j}\to\infty, by (6.12) and (6.13) we have that

EjBrjlocE,|EjBrj|v,as j,\displaystyle E_{j}\cap B_{r_{j}}\stackrel{{\scriptstyle\scriptscriptstyle{{\rm loc}}}}{{\to}}E\,,\qquad|E_{j}\setminus B_{r_{j}}|\to v^{*}\,,\qquad\mbox{as $j\to\infty$}\,, (6.15)
lim infj 2n((Kj)Ej(0)Brj)+n(BrjEj)bk(K,E),\displaystyle\liminf_{j\to\infty}\,2\,\mathcal{H}^{n}\big{(}\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(0)}}\cap B_{r_{j}}\big{)}+\mathcal{H}^{n}(B_{r_{j}}\cap\partial^{*}E_{j})\geq\mathcal{F}_{\rm bk}(K,E)\,, (6.16)

Moreover, since |Ej|<|E_{j}|<\infty, we can choose rjr_{j}\to\infty so that n(Ej(1)Brj)0\mathcal{H}^{n}(E_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{r_{j}})\to 0, while, taking into account that P(EjBrj)=n(Ej(1)Brj)+n((Ej)Brj)P(E_{j}\setminus B_{r_{j}})=\mathcal{H}^{n}(E_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{r_{j}})+\mathcal{H}^{n}((\partial^{*}E_{j})\setminus B_{r_{j}}), we have

bk(Kj,Ej)\displaystyle\mathcal{F}_{\rm bk}(K_{j},E_{j}) \displaystyle\geq 2n((Kj)Ej(0)Brj)+n(BrjEj)\displaystyle 2\,\mathcal{H}^{n}\big{(}\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(0)}}\cap B_{r_{j}}\big{)}+\mathcal{H}^{n}(B_{r_{j}}\cap\partial^{*}E_{j})
+P(EjBrj)n(Ej(1)Brj).\displaystyle+P(E_{j}\setminus B_{r_{j}})-\mathcal{H}^{n}(E_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{r_{j}})\,.

By combining these facts with (6.15), (6.16), and the Euclidean isoperimetric inequality, we conclude that

Ψbk(v)=limjbk(Kj,Ej)bk(K,E)+(n+1)ωn+11/(n+1)limj|EjBrj|n/(n+1),\displaystyle\Psi_{\rm bk}(v)=\lim_{j\to\infty}\mathcal{F}_{\rm bk}(K_{j},E_{j})\geq\mathcal{F}_{\rm bk}(K,E)+(n+1)\,\omega_{n+1}^{1/(n+1)}\,\lim_{j\to\infty}|E_{j}\setminus B_{r_{j}}|^{n/(n+1)}\,,

that is (6.14).

Step four: We prove the existence of minimizers in Ψbk(v)\Psi_{\rm bk}(v), v>0v>0. By step three, there is (K,E)𝒦B(K,E)\in\mathcal{K}_{\rm B} such that KE(1)K\cup E^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, (K,E)(K,E) is a minimizer of Ψbk(|E|)\Psi_{\rm bk}(|E|) and, combining (6.11) and (6.14),

Ψbk(v)=Ψbk(|E|)+(n+1)ωn+11/(n+1)(v|E|)n/(n+1).\Psi_{\rm bk}(v)=\Psi_{\rm bk}(|E|)+(n+1)\,\omega_{n+1}^{1/(n+1)}\,(v-|E|)^{n/(n+1)}\,. (6.17)

Since (K,E)(K,E) is a minimizer in Ψbk(|E|)\Psi_{\rm bk}(|E|), by step one we can assume that KK is n\mathcal{H}^{n}-rectifiable and that both KK and EE are bounded. We can thus find Br(x0)ΩB_{r}(x_{0})\subset\!\subset\Omega such that |Br(x0)|=v|E||B_{r}(x_{0})|=v-|E|, |Br(x0)E|=0|B_{r}(x_{0})\cap E|=0, and n(KBr(x0))=0\mathcal{H}^{n}(K\cap B_{r}(x_{0}))=0. In this way (K,E)=(KBr(x0),EBr(x0))𝒦B(K_{*},E_{*})=(K\cup\partial B_{r}(x_{0}),E\cup B_{r}(x_{0}))\in\mathcal{K}_{\rm B} is trivially 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} and such that |E|=v|E_{*}|=v, and thus is a competitor for Ψbk(v)\Psi_{\rm bk}(v). At the same time,

bk(K,E)=bk(K,E)+(n+1)ωn+11/(n+1)(v|E|)n/(n+1)\mathcal{F}_{\rm bk}(K_{*},E_{*})=\mathcal{F}_{\rm bk}(K,E)+(n+1)\,\omega_{n+1}^{1/(n+1)}\,(v-|E|)^{n/(n+1)}

so that, by (6.17), (K,E)(K_{*},E_{*}) is a minimizer of Ψbk(v)\Psi_{\rm bk}(v). Having proved that minimizers of Ψbk(v)\Psi_{\rm bk}(v) do indeed exist, a further application of step one completes the proof of statement (i).

Step five: We finally prove statement (ii). Let us consider a sequence vj0+v_{j}\to 0^{+} and corresponding minimizers (Kj,Ej)(K_{j},E_{j}) of Ψbk(vj)\Psi_{\rm bk}(v_{j}). By (6.11) with v1=0v_{1}=0 and v2=vjv_{2}=v_{j} we see that {(Kj,Ej)}j\{(K_{j},E_{j})\}_{j} satisfies the assumptions of Theorem 1.4. Since |Ej|=vj0|E_{j}|=v_{j}\to 0, setting μj=n  (ΩEj)+2n  ((Kj)Ej(0))\mu_{j}=\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E_{j})+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(0)}}), the conclusion of Theorem 1.4 is that there are a Radon measure μ\mu in Ω\Omega and a Borel set KK such that KK is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} and μjμ\mu_{j}\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}\mu for a Radon measure μ\mu in Ω\Omega such that μ2n  K\mu\geq 2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}K. Thanks to (6.11) we thus have

2\displaystyle 2\,\ell =\displaystyle= limjΨbk(0)+(n+1)ωn+11/(n+1)vjn/(n+1)lim infjΨbk(vj)\displaystyle\lim_{j\to\infty}\Psi_{\rm bk}(0)+(n+1)\,\omega_{n+1}^{1/(n+1)}\,v_{j}^{n/(n+1)}\geq\liminf_{j\to\infty}\Psi_{\rm bk}(v_{j})
=\displaystyle= lim infjbk(Kj,Ej)bk(K,)=2n(K)2.\displaystyle\liminf_{j\to\infty}\mathcal{F}_{\rm bk}(K_{j},E_{j})\geq\mathcal{F}_{\rm bk}(K,\emptyset)=2\,\mathcal{H}^{n}(K)\geq 2\,\ell\,.

We conclude that Ψbk(vj)2\Psi_{\rm bk}(v_{j})\to 2\,\ell, KK is a minimizer of \ell, and μ=2n  K\mu=2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}K, thus completing the proof of the theorem. ∎

Proof of Theorem 1.5.

The identity (1.22) is proved in Theorem 6.1. Conclusions (i), (ii), and (iii) are proved in Theorem 6.2. ∎

7. Equilibrium across transition lines in wet soap films (Theorem 1.6)

We finally prove Theorem 1.6. We shall need two preliminary lemmas:

Lemma 7.1 (Representation of bk\mathcal{F}_{\rm bk} via induced partitions).

If UΩU\subset\Omega is a set of finite perimeter, (K,E)𝒦B(K,E)\in\mathcal{K}_{\rm B} is such that bk(K,E)<\mathcal{F}_{\rm bk}(K,E)<\infty, and {Ui}i\{U_{i}\}_{i} is a Lebesgue partition of UEU\setminus E induced by KK, then each UiU_{i} has finite perimeter, and, setting K=iUiK^{*}=\bigcup_{i}\partial^{*}U_{i}, we have

bk(K,E;U(1))=in(U(1)Ui)+2n(U(1)(KK)E(0));\displaystyle\mathcal{F}_{\rm bk}(K,E;U^{\scriptscriptstyle{(1)}})=\sum_{i}\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}U_{i})+2\,\mathcal{H}^{n}\big{(}U^{\scriptscriptstyle{(1)}}\cap(K\setminus K^{*})\cap E^{\scriptscriptstyle{(0)}}\big{)}\,; (7.1)

see

Refer to caption
C1C_{1}UUC2C_{2}EEKKKKK\setminus K^{*}(b)(b)(a)(a)
Figure 7.1. The situation in Lemma 7.1: (a) a depiction of the left hand side of (7.1), where KEK\setminus\partial^{*}E is drawn with a bold line to indicate that, in the computation of bk(K,E;U(1))=n(U(1)E)+2n(U(1)KE)\mathcal{F}_{\rm bk}(K,E;U^{\scriptscriptstyle{(1)}})=\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}E)+2\,\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap K\setminus\partial^{*}E), it is counted with multiplicity 22; (b) a depiction of the right hand side of (7.1), where KKK\setminus K^{*} is drawn with a bold line to indicate that it has to be counted with multiplicity 22.

Figure 7.1.

Proof.

For each ii, eUi\partial^{e}U_{i} is contained in (eU)(eE)(UE)(1)(\partial^{e}U)\cup(\partial^{e}E)\cup(U\setminus E)^{\scriptscriptstyle{(1)}}, where both eU\partial^{e}U and eE\partial^{e}E are n\mathcal{H}^{n}-finite being UU and EE of finite perimeter, and where (UE)(1)eUi(U\setminus E)^{\scriptscriptstyle{(1)}}\cap\partial^{e}U_{i} is n\mathcal{H}^{n}-contained in KK by assumption. Now, (UE)(1)n+1E(1)(U\setminus E)^{\scriptscriptstyle{(1)}}\subset\mathbb{R}^{n+1}\setminus E^{\scriptscriptstyle{(1)}}, so that

n((UE)(1)eUi)n(KE(1))bk(K,E)<.\mathcal{H}^{n}\big{(}(U\setminus E)^{\scriptscriptstyle{(1)}}\cap\partial^{e}U_{i}\big{)}\leq\mathcal{H}^{n}(K\setminus E^{\scriptscriptstyle{(1)}})\leq\mathcal{F}_{\rm bk}(K,E)<\infty\,.

This shows that, for each ii, UiU_{i} is a set of finite perimeter. As a consequence {UE}{Ui}i\{U\cap E\}\cup\{U_{i}\}_{i} is a Caccioppoli partition of UU, so that, by (1.46),

2n(U(1)[(UE)K])=n(U(1)(UE))+in(U(1)Ui),2\,\mathcal{H}^{n}\Big{(}U^{\scriptscriptstyle{(1)}}\cap\Big{[}\partial^{*}(U\cap E)\cup K^{*}\Big{]}\Big{)}=\mathcal{H}^{n}\big{(}U^{\scriptscriptstyle{(1)}}\cap\partial^{*}(U\cap E)\big{)}+\sum_{i}\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}U_{i})\,, (7.2)

with K=iUiK^{*}=\bigcup_{i}\partial^{*}U_{i}. Now, thanks to (1.40), (1.41), and the inclusion in (1.46), we have

U(1)(UE)=nU(1)EnU(1)K,U^{\scriptscriptstyle{(1)}}\cap\partial^{*}(U\cap E)\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}U^{\scriptscriptstyle{(1)}}\cap\partial^{*}E\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}U^{\scriptscriptstyle{(1)}}\cap K^{*}\,,

which combined with (7.2) gives

2n(U(1)K)=n(U(1)E)+in(U(1)Ui).2\,\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap K^{*})=\mathcal{H}^{n}\big{(}U^{\scriptscriptstyle{(1)}}\cap\partial^{*}E\big{)}+\sum_{i}\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}U_{i})\,. (7.3)

Therefore, using in order

U(1)EnU(1)K,KnK,n(KE(1))=0,U^{\scriptscriptstyle{(1)}}\cap\partial^{*}E\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}U^{\scriptscriptstyle{(1)}}\cap K^{*}\,,\qquad K^{*}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}K\,,\qquad\mathcal{H}^{n}(K^{*}\cap E^{\scriptscriptstyle{(1)}})=0\,,

and Federer’s theorem (1.37), we obtain

bk(K,E;U(1))\displaystyle\mathcal{F}_{\rm bk}(K,E;U^{\scriptscriptstyle{(1)}}) =\displaystyle= n(U(1)E)+2n(U(1)KE(0))\displaystyle\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}E)+2\,\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap K\cap E^{\scriptscriptstyle{(0)}})
=\displaystyle= 2n(U(1)KE)n(U(1)E)\displaystyle 2\,\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap K^{*}\cap\partial^{*}E)-\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}E)
+2n(U(1)KE(0))+2n(U(1)(KK)E(0))\displaystyle+2\,\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap K^{*}\cap E^{\scriptscriptstyle{(0)}})+2\,\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap(K\setminus K^{*})\cap E^{\scriptscriptstyle{(0)}})
=\displaystyle= 2n(U(1)K)n(U(1)E)+2n(U(1)(KK)E(0))\displaystyle 2\,\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap K^{*})-\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}E)+2\,\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap(K\setminus K^{*})\cap E^{\scriptscriptstyle{(0)}})
=\displaystyle= in(U(1)Ui)+2n(U(1)(KK)E(0)),\displaystyle\sum_{i}\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap\partial^{*}U_{i})+2\,\mathcal{H}^{n}(U^{\scriptscriptstyle{(1)}}\cap(K\setminus K^{*})\cap E^{\scriptscriptstyle{(0)}})\,,

where in the last identity we have used (7.3). ∎

The next lemma is a slight reformulation of [DLGM17a, Lemma 10] and [DLDRG19, Lemma 4.1].

Lemma 7.2.

If 𝐖\mathbf{W} is closed, 𝒞\mathcal{C} is a spanning class for 𝐖\mathbf{W}, SS is relatively closed in Ω\Omega and 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, and BΩB\subset\Omega is an open ball, then for any γ𝒞\gamma\in\mathcal{C} we either have γ(SS1)(SB)\gamma(\SS^{1})\cap(S\setminus B)\neq\varnothing, or γ(SS1)\gamma(\SS^{1}) has non-empty intersection with at least two connected components of BSB\setminus S. In particular, it intersects the boundaries of both components.

We are now ready for the proof of Theorem 1.6.

Proof of Theorem 1.6.

The opening part of the statement of Theorem 1.6 is Theorem 6.2-(i), therefore we can directly consider a minimizer (K,E)𝒦(K,E)\in\mathcal{K} of Ψbk(v)\Psi_{\rm bk}(v) such that both EE and KK are bounded, KEK\cup E is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, and

KE(1)=,K\cap E^{\scriptscriptstyle{(1)}}=\varnothing\,, (7.4)

and begin by proving the existence of a closed set ΣK\Sigma\subset K closed such that (i): Σ=\Sigma=\varnothing if 1n61\leq n\leq 6, Σ\Sigma is locally finite in Ω\Omega if n=7n=7, and s(Σ)=0\mathcal{H}^{s}(\Sigma)=0 for every s>n7s>n-7 if n8n\geq 8; (ii): (E)Σ(\partial^{*}E)\setminus\Sigma is a smooth hypersurface with constant mean curvature; (iii) K(cl(E)Σ)K\setminus(\mathrm{cl}\,(E)\cup\Sigma) is a smooth minimal hypersurface; (iv)α: if x[Ω(EE)]Σx\in[\Omega\cap(\partial E\setminus\partial^{*}E)]\setminus\Sigma, then there are r>0r>0, νSSn\nu\in\SS^{n}, u1,u2C1,α(𝐃rν(x);(r/4,r/4))u_{1},u_{2}\in C^{1,\alpha}(\mathbf{D}_{r}^{\nu}(x);(-r/4,r/4)) (α(0,1/2)\alpha\in(0,1/2) arbitrary) such that u1(x)=u2(x)=0u_{1}(x)=u_{2}(x)=0, u1u2u_{1}\leq u_{2} on 𝐃rν(x)\mathbf{D}_{r}^{\nu}(x), {u1<u2}\{u_{1}<u_{2}\} and int{u1=u2}{\rm int}\{u_{1}=u_{2}\} are both non-empty, and

𝐂rν(x)K\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap K =\displaystyle= i=1,2{y+ui(y)ν:y𝐃rν(x)},\displaystyle\cup_{i=1,2}\big{\{}y+u_{i}(y)\,\nu:y\in\mathbf{D}_{r}^{\nu}(x)\big{\}}\,, (7.5)
𝐂rν(x)E\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap\partial^{*}E =\displaystyle= i=1,2{y+ui(y)ν:y{u1<u2}},\displaystyle\cup_{i=1,2}\big{\{}y+u_{i}(y)\nu:y\in\{u_{1}<u_{2}\}\big{\}}\,, (7.6)
𝐂rν(x)E\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap E =\displaystyle= {y+tν:y{u1<u2},u1(x)<t<u2(x)}.\displaystyle\big{\{}y+t\,\nu:y\in\{u_{1}<u_{2}\}\,,u_{1}(x)<t<u_{2}(x)\big{\}}\,. (7.7)

(The sharp version of conclusion (iv), that is conclusion (iv)α with α=1\alpha=1, and conclusion (v), will be proved in the final step five of this proof.) The key step to prove conclusions (i)–(iv)α is showing the validity of the following claim.

Claim: There exist positive constants Λ\Lambda and r0r_{0} such that if B2r(x)ΩB_{2r}(x)\subset\!\subset\Omega, then, denoting by {Uj}j\{U_{j}\}_{j} the open connected components of B2r(x)(EK)B_{2r}(x)\setminus(E\cup K),

Br(x)K=Br(x)jUj,\displaystyle B_{r}(x)\cap K=B_{r}(x)\cap\cup_{j}\partial U_{j}\,, (7.8)
#{i:Br(x)Uj}<,\displaystyle\#\big{\{}i:B_{r}(x)\cap U_{j}\neq\varnothing\}<\infty\,, (7.9)
B2r(x)cl(Uj)=B2r(x)Uj,\displaystyle B_{2\,r}(x)\cap\mathrm{cl}\,(\partial^{*}U_{j})=B_{2\,r}(x)\cap\partial U_{j}\,, (7.10)
P(Uj;Br(x))P(Vj;Br(x))+Λ|UjΔVj|,\displaystyle P(U_{j};B_{r}(x))\leq P(V_{j};B_{r}(x))+\Lambda\,|U_{j}\Delta V_{j}|\,, (7.11)

whenever VjV_{j} satisfies VjΔUjBr(x)V_{j}\Delta U_{j}\subset\!\subset B_{r}(x) and diam(UjΔVj)<r0{\rm diam}\,(U_{j}\Delta V_{j})<r_{0}.

Deduction of (i)-(iv) from the claim: Let {B2ri(xi)}i\{B_{2r_{i}}(x_{i})\}_{i\in\mathbb{N}} be a countable family of balls, locally finite in Ω\Omega, such that B2ri(xi)ΩB_{2r_{i}}(x_{i})\subset\!\subset\Omega and Ω=iBri(xi)\Omega=\cup_{i}B_{r_{i}}(x_{i}). Setting for brevity

Ωi=Bri(xi),\Omega_{i}=B_{r_{i}}(x_{i})\,,

by (7.9) there are finitely many connected components {Uji}j=1Ji\{U_{j}^{i}\}_{j=1}^{J_{i}} of B2ri(xi)(EK)B_{2r_{i}}(x_{i})\setminus(E\cup K) such that UjiΩiU_{j}^{i}\cap\Omega_{i}\neq\varnothing. Thanks to (7.11), we deduce from [Mag12, Theorem 28.1] that, if we set Σji=Ωi(UjiUji)\Sigma_{j}^{i}=\Omega_{i}\cap(\partial U_{j}^{i}\setminus\partial^{*}U_{j}^{i}), then ΩiUji\Omega_{i}\cap\partial^{*}U_{j}^{i} is a C1,αC^{1,\alpha}-hypersurface for every α(0,1/2)\alpha\in(0,1/2), and Σji\Sigma_{j}^{i} is a closed set that satisfies the dimensional estimates listed in conclusion (i). In particular, if we set

Σ=ij=1JiΣji,\Sigma=\cup_{i\in\mathbb{N}}\cup_{j=1}^{J_{i}}\Sigma^{i}_{j}\,, (7.12)

then ΣK\Sigma\subset K thanks to ΣjiΩiUji\Sigma_{j}^{i}\subset\Omega_{i}\cap\partial U_{j}^{i} and to (7.8), and conclusion (i) holds by the local finiteness of the covering {B2ri(xi)}i\{B_{2r_{i}}(x_{i})\}_{i} of Ω\Omega and from Ji<J_{i}<\infty for every ii. Before moving to prove the remaining conclusions, we first notice that (7.8) gives

ΩiKΣ\displaystyle\Omega_{i}\cap K\setminus\Sigma =\displaystyle= Ωij=1JiUjiΣ\displaystyle\Omega_{i}\cap\cup_{j=1}^{J_{i}}\partial U_{j}^{i}\setminus\Sigma (7.13)
\displaystyle\subset Ωij=1Ji(UjiΣji)=Ωij=1JiUji;\displaystyle\Omega_{i}\cap\cup_{j=1}^{J_{i}}(\partial U_{j}^{i}\setminus\Sigma^{i}_{j})\,\,=\,\,\Omega_{i}\cap\cup_{j=1}^{J_{i}}\partial^{*}U_{j}^{i}\,;

second, we notice that, since KK is n\mathcal{H}^{n}-finite,

{EΩi,UijΩi}j=1Ji is a Caccioppoli partition of Ωi;\mbox{$\{E\cap\Omega_{i},U_{i}^{j}\cap\Omega_{i}\}_{j=1}^{J_{i}}$ is a Caccioppoli partition of $\Omega_{i}$}\,; (7.14)

finally, we recall that, by (1.23), for every XCc1(Ω;n+1)X\in C^{1}_{c}(\Omega;\mathbb{R}^{n+1}) it holds

λEXνE𝑑n=EdivKX𝑑n+2KE(0)divKX𝑑n.\displaystyle\lambda\,\int_{\partial^{*}E}X\cdot\nu_{E}\,d\mathcal{H}^{n}=\int_{\partial^{*}E}\mathrm{div}^{K}\,X\,d\mathcal{H}^{n}+2\,\int_{K\cap E^{{\scriptscriptstyle{(0)}}}}\mathrm{div}^{K}\,X\,d\mathcal{H}^{n}\,. (7.15)

To prove conclusion (ii): Given xΩEΣx\in\Omega\cap\partial^{*}E\setminus\Sigma, there is ii\in\mathbb{N} such that xΩiEx\in\Omega_{i}\cap\partial^{*}E. By ΩEK\Omega\cap\partial^{*}E\subset K and by (7.13) there is j(x){1,,Ji}j(x)\in\{1,...,J_{i}\} such that xUj(x)ix\in\partial^{*}U_{j(x)}^{i}. By (7.14), we can use (1.47) and xΩEUj(x)ix\in\Omega\cap\partial^{*}E\cap\partial^{*}U_{j(x)}^{i} to deduce that

xjj(x)Uji.x\not\in\cup_{j\neq j(x)}\partial^{*}U_{j}^{i}\,. (7.16)

Let r>0r>0 be such that Br(x)Uj(x)iB_{r}(x)\cap\partial^{*}U_{j(x)}^{i} is a C1C^{1}-hypersurface. Since Σ\Sigma contains jUji\cup_{j}\partial U_{j}^{i} and (7.10) holds, (7.16) implies that there is r>0r>0 such that

Br(x)ΩiΣ,Br(x)jUji=Br(x)Uj(x)i=Br(x)Uj(x)i.B_{r}(x)\subset\!\subset\Omega_{i}\setminus\Sigma\,,\qquad B_{r}(x)\cap\cup_{j}\partial U_{j}^{i}=B_{r}(x)\cap\partial U_{j(x)}^{i}=B_{r}(x)\cap\partial^{*}U_{j(x)}^{i}\,. (7.17)

Since Br(x)jj(x)Uji=B_{r}(x)\cap\cup_{j\neq j(x)}\partial U_{j}^{i}=\varnothing and Br(x)Uj(x)iB_{r}(x)\cap U_{j(x)}^{i}\neq\varnothing, we also have that

Br(x)jUji=Br(x)Uj(x)i,B_{r}(x)\cap\cup_{j}U_{j}^{i}=B_{r}(x)\cap U_{j(x)}^{i}\,,

and thus, by (7.14), that {EBr(x),Uj(x)iBr(x)}\{E\cap B_{r}(x),U_{j(x)}^{i}\cap B_{r}(x)\} is an n\mathcal{H}^{n}-partition of Br(x)B_{r}(x). In particular, Br(x)E=Br(x)Uj(x)iB_{r}(x)\cap\partial^{*}E=B_{r}(x)\cap\partial^{*}U_{j(x)}^{i}: intersecting with Br(x)B_{r}(x) in (7.13) and taking into account (7.17), we conclude that

Br(x)K\displaystyle B_{r}(x)\cap K =\displaystyle= Br(x)[ΩiKΣ]Br(x)[Ωij=1JiUji]=Br(x)Uj(x)i\displaystyle B_{r}(x)\cap[\Omega_{i}\cap K\setminus\Sigma]\,\,\subset\,\,B_{r}(x)\cap[\Omega_{i}\cap\cup_{j=1}^{J_{i}}\partial^{*}U_{j}^{i}]\,\,=\,\,B_{r}(x)\cap\partial^{*}U_{j(x)}^{i}
=\displaystyle= Br(x)E,\displaystyle B_{r}(x)\cap\partial^{*}E\,,

and (7.15) implies that, for every XCc1(Br(x);n+1)X\in C^{1}_{c}(B_{r}(x);\mathbb{R}^{n+1}),

λEXνE𝑑n=EdivKX𝑑n.\displaystyle\lambda\int_{\partial^{*}E}X\cdot\nu_{E}\,d\mathcal{H}^{n}=\int_{\partial^{*}E}\mathrm{div}^{K}\,X\,d\mathcal{H}^{n}\,. (7.18)

Hence, E\partial^{*}E can be represented, locally in Br(x)B_{r}(x), as the graph of distributional solutions of class C1,αC^{1,\alpha} to the constant mean curvature equation. By Schauder’s theory, Br(x)EB_{r}(x)\cap\partial^{*}E is a smooth hypersurface whose mean curvature with respect to νE\nu_{E} is equal to λ\lambda thanks to (7.18).

To prove conclusions (iii) and (iv): Let us now pick xK(ΣE)x\in K\setminus(\Sigma\cup\partial^{*}E) and let ii\in\mathbb{N} be such that xΩiKx\in\Omega_{i}\cap K. Let ii\in\mathbb{N} be such that xΩix\in\Omega_{i}. By (7.13) there is j(x){1,,Ji}j(x)\in\{1,...,J_{i}\} such that xUj(x)ix\in\partial^{*}U_{j(x)}^{i}. By (7.14) and by (1.47), either xEx\in\partial^{*}E (which is excluded from the onset), or there is k(x)j(x)k(x)\neq j(x) such that xUk(x)ix\in\partial^{*}U_{k(x)}^{i}. We have thus proved that

xUj(x)iUk(x)i,xjj(x),k(x)Uji.x\in\partial^{*}U_{j(x)}^{i}\cap\partial^{*}U_{k(x)}^{i}\,,\qquad x\not\in\cup_{j\neq j(x),k(x)}\partial^{*}U_{j}^{i}\,. (7.19)

To prove conclusion (iii) we notice that if we are in the case when xK(ΣE)=K(Σcl(E))x\in K\setminus(\Sigma\cup\partial E)=K\setminus(\Sigma\cup\mathrm{cl}\,(E)) (thanks to KE=K\cap E=\varnothing), then xcl(E)x\not\in\mathrm{cl}\,(E) implies that, for some r>0r>0, Br(x)(Σcl(E))=B_{r}(x)\cap(\Sigma\cup\mathrm{cl}\,(E))=\emptyset. In particular, by (7.14) and (7.19), {Br(x)Uj(x)i,Br(x)Uk(x)i}\{B_{r}(x)\cap U_{j(x)}^{i},B_{r}(x)\cap U_{k(x)}^{i}\} is an n\mathcal{H}^{n}-partition of Br(x)B_{r}(x), and by (7.13)

Br(x)K=Br(x)Uj(x)i=Br(x)Uk(x)i,B_{r}(x)\cap K=B_{r}(x)\cap\partial^{*}U_{j(x)}^{i}=B_{r}(x)\cap\partial^{*}U_{k(x)}^{i}\,,

is a C1,αC^{1,\alpha}-hypersurface. Under these conditions, (7.15) boils down to

KdivKX𝑑n=0,XCc1(Br(x);n+1),\displaystyle\int_{K}\mathrm{div}^{K}\,X\,d\mathcal{H}^{n}=0\,,\qquad\forall X\in C^{1}_{c}(B_{r}(x);\mathbb{R}^{n+1})\,, (7.20)

so that KK can be represented, locally in Br(x)B_{r}(x), as the graph of distributional solutions to the minimal surfaces equation of class C1,αC^{1,\alpha}. By Schauder’s theory, Br(x)KB_{r}(x)\cap K is a smooth minimal surface.

To finally prove conclusion (iv), let us assume that xΩ(EE)Σx\in\Omega\cap(\partial E\setminus\partial^{*}E)\setminus\Sigma. In this case (7.14) and (7.19) do not imply that {Br(x)Uj(x)i,Br(x)Uk(x)i}\{B_{r}(x)\cap U_{j(x)}^{i},B_{r}(x)\cap U_{k(x)}^{i}\} is an n\mathcal{H}^{n}-partition of Br(x)B_{r}(x); actually, by ΩE=Ωcl(E)\Omega\cap\partial E=\Omega\cap\mathrm{cl}\,(\partial^{*}E), the fact that xEx\in\partial E implies that Bs(x)EB_{s}(x)\cap\partial^{*}E\neq\emptyset for every s>0s>0, so that |Bs(x)E|>0|B_{s}(x)\cap E|>0 for every s>0s>0, and the situation is such that, for every s<rs<r,

{Bs(x)Uj(x)i,Bs(x)Uk(x)i,Bs(x)E}\{B_{s}(x)\cap U_{j(x)}^{i},B_{s}(x)\cap U_{k(x)}^{i},B_{s}(x)\cap E\} is an n\mathcal{H}^{n}-partition of Bs(x)B_{s}(x) (7.21)

with all three sets in the partition having positive measure.

Now, by the first inclusion in (7.19), there exists νSSn\nu\in\SS^{n} such that, up to further decrease the value of rr and for some u1,u2C1,α(𝐃rν(x);(r/4,r/4))u_{1},u_{2}\in C^{1,\alpha}(\mathbf{D}_{r}^{\nu}(x);(-r/4,r/4)) with u1(x)=u2(x)=0u_{1}(x)=u_{2}(x)=0 and u1(x)=u2(x)=0\nabla u_{1}(x)=\nabla u_{2}(x)=0 it must hold

𝐂rν(x)Uj(x)i={y+tν:y𝐃rν(x),t>u2(y)},\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap U_{j(x)}^{i}=\big{\{}y+t\,\nu:y\in\mathbf{D}_{r}^{\nu}(x)\,,t>u_{2}(y)\big{\}}\,,
𝐂rν(x)Uk(x)i={y+tν:y𝐃rν(x),t<u1(y)}.\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap U_{k(x)}^{i}=\big{\{}y+t\,\nu:y\in\mathbf{D}_{r}^{\nu}(x)\,,t<u_{1}(y)\big{\}}\,.

By Uj(x)iUk(x)i=U_{j(x)}^{i}\cap U_{k(x)}^{i}=\varnothing we have u1u2u_{1}\leq u_{2} on 𝐃rν(x)\mathbf{D}_{r}^{\nu}(x), so that (7.21) gives

𝐂rν(x)E={y+tν:y{u1<u2},u1(y)<t<u2(y)},\mathbf{C}_{r}^{\nu}(x)\cap E=\big{\{}y+t\,\nu:y\in\{u_{1}<u_{2}\}\,,u_{1}(y)<t<u_{2}(y)\big{\}}\,,

and {u1<u2}\{u_{1}<u_{2}\} is non-empty. Again by (7.19) and (7.13) we also have that

𝐂rν(x)K\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap K =\displaystyle= k=12{y+uk(y)ν:y𝐃rν(x)},\displaystyle\cup_{k=1}^{2}\,\big{\{}y+u_{k}(y)\,\nu:y\in\mathbf{D}_{r}^{\nu}(x)\big{\}}\,,
𝐂rν(x)Uj(x)iUk(x)i\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap\partial^{*}U_{j(x)}^{i}\cap\partial^{*}U_{k(x)}^{i} =\displaystyle= {y+u1(y)ν:y𝐃rν(x){u1=u2}},\displaystyle\big{\{}y+u_{1}(y)\,\nu:y\in\mathbf{D}_{r}^{\nu}(x)\cap\{u_{1}=u_{2}\}\big{\}}\,,
𝐂rν(x)E\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap\partial^{*}E =\displaystyle= k=12{y+uk(y)ν:y𝐃rν(x){u1<u2}}.\displaystyle\cup_{k=1}^{2}\,\big{\{}y+u_{k}(y)\,\nu:y\in\mathbf{D}_{r}^{\nu}(x)\cap\{u_{1}<u_{2}\}\big{\}}\,.

This completes the proof of conclusion (iv)α.

Proof of the claim: Assuming without loss of generality that x=0x=0, we want to find Λ\Lambda and r0r_{0} positive such that if B2rΩB_{2r}\subset\!\subset\Omega, then, denoting by {Uj}j\{U_{j}\}_{j} the open connected components of B2r(EK)B_{2r}\setminus(E\cup K), we have

BrK=BrjUj,\displaystyle B_{r}\cap K=B_{r}\cap\cup_{j}\partial U_{j}\,, (7.22)
#{j:BrUj}<,\displaystyle\#\big{\{}j:B_{r}\cap U_{j}\neq\varnothing\big{\}}<\infty\,, (7.23)
B2rcl(Uj)=B2rUj,\displaystyle B_{2\,r}\cap\mathrm{cl}\,(\partial^{*}U_{j})=B_{2\,r}\cap\partial U_{j}\,, (7.24)

and that P(Uj;Br)P(Vj;Br)+Λ|UjΔVj|P(U_{j};B_{r})\leq P(V_{j};B_{r})+\Lambda\,|U_{j}\Delta V_{j}| whenever VjV_{j} satisfies VjΔUjBrV_{j}\Delta U_{j}\subset\!\subset B_{r} and diam(UjΔVj)<r0{\rm diam}\,(U_{j}\Delta V_{j})<r_{0}.

Step one: We prove that

KintUj(1)=,intUj(1)=Ujj.\displaystyle K\cap\mathrm{int}\,U_{j}^{\scriptscriptstyle{(1)}}=\varnothing\,,\qquad\mathrm{int}\,U_{j}^{\scriptscriptstyle{(1)}}=U_{j}\quad\forall j\,. (7.25)

To this end, we begin by noticing that, for every jj,

B2rUj\displaystyle B_{2\,r}\cap\partial U_{j} \displaystyle\subset B2rK,\displaystyle B_{2\,r}\cap K\,, (7.26)
Ujint(Uj(1))\displaystyle U_{j}\,\,\subset\,\,{\rm int}(U_{j}^{\scriptscriptstyle{(1)}}) \displaystyle\subset B2rclUjB2r(UjK),\displaystyle B_{2\,r}\cap\mathrm{cl}\,U_{j}\,\,\subset\,\,B_{2\,r}\cap(U_{j}\cup K)\,, (7.27)
B2r[int(Uj(1))]\displaystyle B_{2\,r}\cap\partial[{\rm int}(U_{j}^{\scriptscriptstyle{(1)}})] \displaystyle\subset B2rK.\displaystyle B_{2\,r}\cap K\,. (7.28)

Indeed, for every kk and jj, UkUj=U_{k}\cap U_{j}=\varnothing with UkU_{k} and UjU_{j} open gives UkUj=U_{k}\cap\partial U_{j}=\varnothing, so that B2rUjB2rkUk=B2r(EK)=B2rKB_{2r}\cap\partial U_{j}\subset B_{2r}\setminus\cup_{k}U_{k}=B_{2\,r}\cap(E\cup K)=B_{2\,r}\cap K thanks to the fact that EUj=E\cap\partial U_{j}=\varnothing (as UjE=U_{j}\cap E=\varnothing). Having proved (7.26), one easily deduces the third inclusion in (7.27), while the first two are evident. Finally, from (7.27), and since KK is closed, we find

B2rcl(int(Uj(1)))B2r(cl(Uj)K),B_{2\,r}\cap\mathrm{cl}\,\big{(}{\rm int}(U_{j}^{\scriptscriptstyle{(1)}})\big{)}\subset B_{2\,r}\cap(\mathrm{cl}\,(U_{j})\cup K)\,,

so that subtracting int(Uj(1)){\rm int}(U_{j}^{\scriptscriptstyle{(1)}}), and recalling that Ujint(Uj(1))U_{j}\subset{\rm int}(U_{j}^{\scriptscriptstyle{(1)}}) we find

B2r[int(Uj(1))]B2r(KUj)B_{2\,r}\cap\partial[{\rm int}(U_{j}^{\scriptscriptstyle{(1)}})]\subset B_{2\,r}\cap(K\cup\partial U_{j})

and deduce (7.28) from (7.26).

Next, we claim that,

if K=KjintUj(1), then (K,E)𝒦 and KE is 𝒞-spanning.\mbox{if $K_{*}=K\setminus\bigcup_{j}\mathrm{int}\,U_{j}^{\scriptscriptstyle{(1)}}$, then $(K_{*},E)\in\mathcal{K}$ and $K_{*}\cup E$ is $\mathcal{C}$-spanning}\,. (7.29)

To prove that (K,E)𝒦(K_{*},E)\in\mathcal{K}, the only assertion that is not immediate is the inclusion ΩEK\Omega\cap\partial E\subset K_{*}. To prove it we notice that if zintUj(1)z\in\mathrm{int}\,U_{j}^{\scriptscriptstyle{(1)}}, then Bs(z)intUj(1)B_{s}(z)\subset\mathrm{int}\,U_{j}^{\scriptscriptstyle{(1)}} for some s>0s>0, so that UjE=U_{j}\cap E=\varnothing gives |EBs(z)|=0|E\cap B_{s}(z)|=0. Since EE is open this implies Bs(z)E=B_{s}(z)\cap E=\varnothing, hence zEz\notin\partial E.

To prove that EKE\cup K_{*} is 𝒞\mathcal{C}-spanning: Since EKE\cup K_{*} is relatively closed in Ω\Omega, it suffices to verify that for arbitrary γ𝒞\gamma\in\mathcal{C}, (KE)γ(K_{*}\cup E)\cap\gamma\neq\varnothing. Since KB2r=KB2rK\setminus B_{2r}=K_{*}\setminus B_{2r}, we directly assume that (KE)(γB2r)=(K\cup E)\cap(\gamma\setminus B_{2r})=\varnothing. Since KEK\cup E is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, by Lemma 7.2, there are two distinct connected components UjU_{j} and UkU_{k} of B2r(KE)B_{2r}\setminus(K\cup E) such that there is γ(SS1)B2r(Uj)(Uk)\gamma(\SS^{1})\cap B_{2\,r}\cap(\partial U_{j})\cap(\partial U_{k})\neq\varnothing. We conclude by showing that

B2r(Uj)(Uk)K,jk.B_{2\,r}\cap(\partial U_{j})\cap(\partial U_{k})\subset K_{*}\,,\qquad\forall j\neq k\,. (7.30)

Indeed any point in B2r(Uj)(Uk)B_{2r}\cap(\partial U_{j})\cap(\partial U_{k}) is an accumulation point for both UjU_{j} and UkU_{k}, and thus, by (7.27), for both intUj(1){\rm int}U_{j}^{\scriptscriptstyle{(1)}} and intUk(1){\rm int}U_{k}^{\scriptscriptstyle{(1)}}. Since UjUk=U_{j}\cap U_{k}=\emptyset implies (intUj(1))(intUk(1))=({\rm int}U_{j}^{\scriptscriptstyle{(1)}})\cap({\rm int}U_{k}^{\scriptscriptstyle{(1)}})=\emptyset, an accumulation point for both intUj(1){\rm int}U_{j}^{\scriptscriptstyle{(1)}} and intUk(1){\rm int}U_{k}^{\scriptscriptstyle{(1)}} must lie in [(intUj(1))][(intUk(1))][\partial({\rm int}U_{j}^{\scriptscriptstyle{(1)}})]\cap[\partial({\rm int}U_{k}^{\scriptscriptstyle{(1)}})]. We thus deduce (7.30) from (7.28), and complete the proof of (7.29).

To deduce (7.25) from (7.29), and complete step one: By (7.29), (K,E)(K_{*},E) is admissible in Ψbk(v)\Psi_{\rm bk}(v). Since (K,E)(K,E) is a minimizer of Ψbk(v)\Psi_{\rm bk}(v), we conclude that n(KK)=0\mathcal{H}^{n}(K\setminus K_{*})=0. Would there be zint(Uj(1))Kz\in{\rm int}(U_{j}^{\scriptscriptstyle{(1)}})\cap K for some jj, then by (6.3), and with ρ>0\rho>0 such that Bρ(z)int(Uj(1))B_{\rho}(z)\subset{\rm int}(U_{j}^{\scriptscriptstyle{(1)}}), we would find

cρnn(KBρ(z))n(Kint(Uj(1)))n(KK)=0.c\,\rho^{n}\leq\mathcal{H}^{n}(K\cap B_{\rho}(z))\leq\mathcal{H}^{n}(K\cap{\rm int}(U_{j}^{\scriptscriptstyle{(1)}}))\leq\mathcal{H}^{n}(K\setminus K_{*})=0\,.

This shows that Kint(Uj(1))=K\cap{\rm int}(U_{j}^{\scriptscriptstyle{(1)}})=\varnothing. Using this last fact in combination with int(Uj(1))B2r(UjK){\rm int}(U_{j}^{\scriptscriptstyle{(1)}})\subset B_{2\,r}\cap(U_{j}\cap K) from (7.27) we conclude that int(Uj(1))Uj{\rm int}(U_{j}^{\scriptscriptstyle{(1)}})\subset U_{j}, and thus that int(Uj(1))=Uj{\rm int}(U_{j}^{\scriptscriptstyle{(1)}})=U_{j} by the first inclusion in (7.27).

Step two: We prove (7.24), i.e. B2rcl(Uj)=B2rUjB_{2\,r}\cap\mathrm{cl}\,(\partial^{*}U_{j})=B_{2\,r}\cap\partial U_{j}. The \subset inclusion is a general fact, see (1.35). To prove the reverse inclusion we recall, again from (1.35), that zB2rcl(Uj)z\in B_{2\,r}\cap\mathrm{cl}\,(\partial^{*}U_{j}) if and only if 0<|Bρ(z)Uj|<|Bρ|0<|B_{\rho}(z)\cap U_{j}|<|B_{\rho}| for every ρ>0\rho>0. Now, if zB2rUjz\in B_{2\,r}\cap\partial U_{j}, then clearly, being UjU_{j} open, we have |UjBρ(z)|>0|U_{j}\cap B_{\rho}(z)|>0 for every ρ>0\rho>0; moreover, should |Bρ(z)Uj|=|Bρ||B_{\rho}(z)\cap U_{j}|=|B_{\rho}| hold for some ρ\rho, then we would have zint(Uj(1))z\in{\rm int}(U_{j}^{\scriptscriptstyle{(1)}}), and thus zUjz\in U_{j} by (7.25), a contradiction.

Step three: We prove, for each jj, the n\mathcal{H}^{n}-equivalence of Uj\partial^{*}U_{j} and Uj\partial U_{j}, that is

n(B2rUjUj)=0.\displaystyle\mathcal{H}^{n}(B_{2\,r}\cap\partial U_{j}\setminus\partial^{*}U_{j})=0\,. (7.31)

By a standard argument [Mag12, Theorem 21.11] it will suffice to prove the existence of r0>0r_{0}>0 and α,β(0,1/2)\alpha,\beta\in(0,1/2) (depending on nn) such that, for each jj and each zB2rUjz\in B_{2\,r}\cap\partial U_{j}, it holds

α|Bρ||Bρ(z)Uj|(1β)|Bρ|,\displaystyle\alpha\,|B_{\rho}|\leq|B_{\rho}(z)\cap U_{j}|\leq(1-\beta)|B_{\rho}|\,, (7.32)

for every ρ<min{r0,dist(z,B2r)}\rho<\min\{r_{0},{\rm dist}(z,\partial B_{2\,r})\}.

Proof of the lower bound in (7.32): Since diffeomorphic images of 𝒞\mathcal{C}-spanning sets are 𝒞\mathcal{C}-spanning, a standard argument using diffeomorphic volume fixing variations shows the existence of positive constants Λ\Lambda and r0r_{0} such that if (K,E)𝒦B(K^{\prime},E^{\prime})\in\mathcal{K}_{\rm B}, K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, and (KΔK)(EΔE)Bρ(z)(K^{\prime}\Delta K)\cup(E^{\prime}\Delta E)\subset\!\subset B_{\rho}(z) for some ρ<r0\rho<r_{0} and Bρ(z)B2rB_{\rho}(z)\subset\!\subset B_{2\,r}, then

bk(K,E)bk(K,E)+Λ|EΔE|.\displaystyle\mathcal{F}_{\rm bk}(K,E)\leq\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime})+\Lambda\,|E\Delta E^{\prime}|\,. (7.33)

We claim that we can apply (7.33) with

E=E(Bρ(z)clUj),K=(K(Uj(1)Bρ(z))(E)(1),\displaystyle E^{\prime}=E\cup\big{(}B_{\rho}(z)\cap\mathrm{cl}\,U_{j}\big{)}\,,\quad K^{\prime}=\big{(}K\cup(U_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z)\big{)}\setminus(E^{\prime})^{\scriptscriptstyle{(1)}}\,, (7.34)

where ρ<r0\rho<r_{0}, Bρ(z)B2rB_{\rho}(z)\subset\!\subset B_{2\,r}, and

n(Bρ(z)[EUj])=n(KBρ(z))=0.\mathcal{H}^{n}\big{(}\partial B_{\rho}(z)\cap[\partial^{*}E\cup\partial^{*}U_{j}]\big{)}=\mathcal{H}^{n}(K\cap\partial B_{\rho}(z))=0\,. (7.35)

Indeed, K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}} contains KE(1)K\cup E^{\scriptscriptstyle{(1)}}, thus KEK\cup E being EE open, and is thus 𝒞\mathcal{C}-spanning. To check that (K,E)𝒦B(K^{\prime},E^{\prime})\in\mathcal{K}_{\rm B}, we argue as follows. First, we notice that n({νE=νBρ(z)cl(Uj)})=0\mathcal{H}^{n}(\{\nu_{E}=\nu_{B_{\rho}(z)\cap\mathrm{cl}\,(U_{j})}\})=0, since it is n\mathcal{H}^{n}-contained in the union of Bρ(z)E\partial B_{\rho}(z)\cap\partial^{*}E and {νE=νcl(Uj)}\{\nu_{E}=\nu_{\mathrm{cl}\,(U_{j})}\}, that are n\mathcal{H}^{n}-negligible by (7.35) and by the fact that νE=νcl(Uj)\nu_{E}=-\nu_{\mathrm{cl}\,(U_{j})} n\mathcal{H}^{n}-a.e. on Ecl(Uj)\partial^{*}E\cap\partial^{*}\mathrm{cl}\,(U_{j}) thanks to |Ecl(Uj)|=0|E\cap\mathrm{cl}\,(U_{j})|=0. By n({νE=νBρ(z)cl(Uj)})=0\mathcal{H}^{n}(\{\nu_{E}=\nu_{B_{\rho}(z)\cap\mathrm{cl}\,(U_{j})}\})=0 and (1.39) we thus have

ΩE=nΩ{[E(0)(Bρ(z)clUj)][(Bρ(z)clUj)(0)E]}.\Omega\cap\partial^{*}E^{\prime}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\Omega\cap\big{\{}\big{[}E^{\scriptscriptstyle{(0)}}\cap\partial^{*}\big{(}B_{\rho}(z)\cap\mathrm{cl}\,U_{j}\big{)}\big{]}\cup\big{[}\big{(}B_{\rho}(z)\cap\mathrm{cl}\,U_{j}\big{)}^{\scriptscriptstyle{(0)}}\cap\partial^{*}E\big{]}\big{\}}\,. (7.36)

Since UjU_{j} is Lebesgue equivalent to cl(Uj)\mathrm{cl}\,(U_{j}) (indeed, B2rUjKB_{2\,r}\cap\partial U_{j}\subset K), we have Uj(1)=[cl(Uj)](1)U_{j}^{\scriptscriptstyle{(1)}}=[\mathrm{cl}\,(U_{j})]^{\scriptscriptstyle{(1)}} and [cl(Uj)]=Uj\partial^{*}[\mathrm{cl}\,(U_{j})]=\partial^{*}U_{j}, so that (1.40) and (7.35) give

(Bρ(z)cl(Uj))=n{[cl(Uj)](1)Bρ(z)}{Bρ(x)[cl(Uj)]},\displaystyle\partial^{*}\big{(}B_{\rho}(z)\cap\mathrm{cl}\,(U_{j})\big{)}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\big{\{}[\mathrm{cl}\,(U_{j})]^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z)\big{\}}\cup\big{\{}B_{\rho}(x)\cap\partial^{*}[\mathrm{cl}\,(U_{j})]\big{\}}\,,
=(Uj(1)Bρ(z))(Bρ(x)Uj)(Uj(1)Bρ(z))K,\displaystyle=\big{(}U_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z)\big{)}\cup\big{(}B_{\rho}(x)\cap\partial^{*}U_{j}\big{)}\subset\big{(}U_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z)\big{)}\cup K\,, (7.37)

by B2rUjKB_{2\,r}\cap\partial U_{j}\subset K. By (7.36) and n((E)(1)E)=0\mathcal{H}^{n}((E^{\prime})^{\scriptscriptstyle{(1)}}\cap\partial^{*}E^{\prime})=0 we thus find that

ΩE(Bρ(z)cl(Uj))nK.\Omega\cap\partial^{*}E^{\prime}\cap\partial^{*}\big{(}B_{\rho}(z)\cap\mathrm{cl}\,(U_{j})\big{)}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}K^{\prime}\,. (7.38)

Moreover, by ΩEΩEK\Omega\cap\partial^{*}E\subset\Omega\cap\partial E\subset K and

(E)(Bρ(z)clUj)(0)E(1/2)(Bρ(z)clUj)(0)n+1(E)(1),(\partial^{*}E)\cap\big{(}B_{\rho}(z)\cap\mathrm{cl}\,U_{j}\big{)}^{\scriptscriptstyle{(0)}}\subset E^{\scriptscriptstyle{(1/2)}}\cap\big{(}B_{\rho}(z)\cap\mathrm{cl}\,U_{j}\big{)}^{\scriptscriptstyle{(0)}}\subset\mathbb{R}^{n+1}\setminus(E^{\prime})^{\scriptscriptstyle{(1)}}\,,

we find (E)(Bρ(z)clUj)(0)K(E)(1)K(\partial^{*}E)\cap\big{(}B_{\rho}(z)\cap\mathrm{cl}\,U_{j}\big{)}^{\scriptscriptstyle{(0)}}\subset K\setminus(E^{\prime})^{\scriptscriptstyle{(1)}}\subset K^{\prime}, which combined with (7.38) finally proves the n\mathcal{H}^{n}-containment of ΩE\Omega\cap\partial^{*}E^{\prime} into KK^{\prime}, and thus (K,E)𝒦B(K^{\prime},E^{\prime})\in\mathcal{K}_{\rm B}. We have thus proved that (K,E)(K^{\prime},E^{\prime}) as in (7.34) is admissible into (7.33). Since bk(K,E;Bρ(z))=0\mathcal{F}_{\rm bk}(K,E;\partial B_{\rho}(z))=0 by (7.35) and bk(K,E;A)=bk(K,E;A)\mathcal{F}_{\rm bk}(K,E;A)=\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime};A) if A=Ωcl(Bρ(z))A=\Omega\setminus\mathrm{cl}\,(B_{\rho}(z)), we deduce from (7.33) that

bk(K,E;Bρ(z))bk(K,E;cl(Bρ(z)))+Λ|EΔE|.\displaystyle\mathcal{F}_{\rm bk}(K,E;B_{\rho}(z))\leq\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime};\mathrm{cl}\,(B_{\rho}(z)))+\Lambda\,|E\Delta E^{\prime}|\,. (7.39)

To exploit (7.39), we first notice that {Bρ(z)Uk}k\{B_{\rho}(z)\cap U_{k}\}_{k} is a Lebesgue partition of Bρ(z)EB_{\rho}(z)\setminus E with Bρ(z)(1)(Bρ(z)Uk)=Bρ(z)UkB_{\rho}(z)^{\scriptscriptstyle{(1)}}\cap\partial^{*}(B_{\rho}(z)\cap U_{k})=B_{\rho}(z)\cap\partial^{*}U_{k} for every kk, so that, by Lemma 7.1,

bk(K,E;Bρ(z))=2n(Bρ(z)E(0)(KkUk))+kP(Uk;Bρ(z)).\mathcal{F}_{\rm bk}(K,E;B_{\rho}(z))=2\,\mathcal{H}^{n}\Big{(}B_{\rho}(z)\cap E^{\scriptscriptstyle{(0)}}\cap\Big{(}K\setminus\bigcup_{k}\partial^{*}U_{k}\Big{)}\Big{)}+\sum_{k}P(U_{k};B_{\rho}(z))\,. (7.40)

Similarly, {Bρ(z)Uk}kj\{B_{\rho}(z)\cap U_{k}\}_{k\neq j} is a Lebesgue partition of Bρ(z)EB_{\rho}(z)\setminus E^{\prime}, so that again by Lemma 7.1 we find

bk(K,E;Bρ(z))=2n(Bρ(z)(E)(0)(KkjUk))+kjP(Uk;Bρ(z))\displaystyle\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime};B_{\rho}(z))=2\,\mathcal{H}^{n}\Big{(}B_{\rho}(z)\cap(E^{\prime})^{\scriptscriptstyle{(0)}}\cap\Big{(}K^{\prime}\setminus\bigcup_{k\neq j}\partial^{*}U_{k}\Big{)}\Big{)}+\sum_{k\neq j}P(U_{k};B_{\rho}(z))
=2n(Bρ(z)(E)(0)(KkUk))+kjP(Uk;Bρ(z))\displaystyle=2\,\mathcal{H}^{n}\Big{(}B_{\rho}(z)\cap(E^{\prime})^{\scriptscriptstyle{(0)}}\cap\Big{(}K\setminus\bigcup_{k}\partial^{*}U_{k}\Big{)}\Big{)}+\sum_{k\neq j}P(U_{k};B_{\rho}(z)) (7.41)

where in the last identity we have used that, by (7.34), we have Bρ(z)(E)(0)Uj=0B_{\rho}(z)\cap(E^{\prime})^{\scriptscriptstyle{(0)}}\cap\partial^{*}U_{j}=0 and Bρ(z)K(E)(0)=Bρ(z)K(E)(0)B_{\rho}(z)\cap K^{\prime}\cap(E^{\prime})^{\scriptscriptstyle{(0)}}=B_{\rho}(z)\cap K\cap(E^{\prime})^{\scriptscriptstyle{(0)}}. Combining (7.39), (7.40), (7.41) and the fact that (E)(0)E(0)(E^{\prime})^{\scriptscriptstyle{(0)}}\subset E^{\scriptscriptstyle{(0)}}, we find that

P(Uj;Bρ(z))bk(K,E;Bρ(z))+Λ|Bρ(z)Uj|.\displaystyle P(U_{j};B_{\rho}(z))\leq\mathcal{F}_{\rm bk}\big{(}K^{\prime},E^{\prime};\partial B_{\rho}(z)\big{)}+\Lambda\,|B_{\rho}(z)\cap U_{j}|\,. (7.42)

The first term in bk(K,E;Bρ(z))\mathcal{F}_{\rm bk}\big{(}K^{\prime},E^{\prime};\partial B_{\rho}(z)\big{)} is P(E;Bρ(z))P(E^{\prime};\partial B_{\rho}(z)): taking into account n(EBρ(z))=0\mathcal{H}^{n}(\partial^{*}E\cap\partial B_{\rho}(z))=0, by (7.36) and the second identity in (7.37) we find

P(E;Bρ(z))\displaystyle P(E^{\prime};\partial B_{\rho}(z)) =\displaystyle= n(Bρ(z)E(0)(Bρ(z)clUj))\displaystyle\mathcal{H}^{n}\big{(}\partial B_{\rho}(z)\cap E^{\scriptscriptstyle{(0)}}\cap\partial^{*}\big{(}B_{\rho}(z)\cap\mathrm{cl}\,U_{j}\big{)}\big{)}
=\displaystyle= n(E(0)Uj(1)Bρ(z))=n(Uj(1)Bρ(z)),\displaystyle\mathcal{H}^{n}(E^{\scriptscriptstyle{(0)}}\cap U_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z))=\mathcal{H}^{n}(U_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z))\,,

while for the second term in bk(K,E;Bρ(z))\mathcal{F}_{\rm bk}\big{(}K^{\prime},E^{\prime};\partial B_{\rho}(z)\big{)}, by n(KBρ(z))=0\mathcal{H}^{n}(K\cap\partial B_{\rho}(z))=0,

n(K(E)(0)Bρ(z))=n((E)(0)Uj(1)Bρ(z))=0\mathcal{H}^{n}(K^{\prime}\cap(E^{\prime})^{\scriptscriptstyle{(0)}}\cap\partial B_{\rho}(z))=\mathcal{H}^{n}((E^{\prime})^{\scriptscriptstyle{(0)}}\cap U_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z))=0

since (E)(0)(Bρ(z)cl(Uj))(0)(E^{\prime})^{\scriptscriptstyle{(0)}}\subset(B_{\rho}(z)\cap\mathrm{cl}\,(U_{j}))^{\scriptscriptstyle{(0)}} and Bρ(z)cl(Uj)B_{\rho}(z)\cap\mathrm{cl}\,(U_{j}) has positive Lebesgue density at points in Uj(1)Bρ(z)U_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z). Having thus proved that bk(K,E;Bρ(z))=n(Uj(1)Bρ(z))\mathcal{F}_{\rm bk}\big{(}K^{\prime},E^{\prime};\partial B_{\rho}(z)\big{)}=\mathcal{H}^{n}(U_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z)), we conclude from (7.42) that

P(Uj;Bρ(z))n(Uj(1)Bρ(z))+Λ|Bρ(z)Uj|,P(U_{j};B_{\rho}(z))\leq\mathcal{H}^{n}(U_{j}^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho}(z))+\Lambda\,|B_{\rho}(z)\cap U_{j}|\,,

for a.e. ρ<r0\rho<r_{0}. Since zB2rUj=B2rcl(Uj)z\in B_{2\,r}\cap\partial U_{j}=B_{2\,r}\cap\mathrm{cl}\,(\partial^{*}U_{j}) and (1.35) imply that |Bρ(z)Uj|>0|B_{\rho}(z)\cap U_{j}|>0 for every ρ>0\rho>0, a standard argument (see, e.g. [Mag12, Theorem 21.11]) implies that, up to further decrease the value of r0r_{0} depending on Λ\Lambda, and for some constant α=α(n)(0,1/2)\alpha=\alpha(n)\in(0,1/2), the lower bound in (7.32) holds true.

Proof of the upper bound in (7.32): We argue by contradiction that, no matter how small β(0,1/2)\beta\in(0,1/2) is, we can find jj, zB2rUjz\in B_{2\,r}\cap\partial U_{j}, and ρ<min{r0,dist(z,B2r)}\rho<\min\{r_{0},{\rm dist}(z,\partial B_{2\,r})\}, such that

|Bρ(z)Uj|>(1β)|Bρ|.\displaystyle|B_{\rho}(z)\cap U_{j}|>(1-\beta)\,|B_{\rho}|\,. (7.43)

We first notice that for every kjk\neq j it must be Bρ/2(z)Uk=B_{\rho/2}(z)\cap\partial U_{k}=\varnothing: indeed if wBρ/2(z)Ukw\in B_{\rho/2}(z)\cap\partial U_{k} for some kjk\neq j, then by the lower bound in (6.2) and by (7.43) we find

α|Bρ/2||UkBρ/2(w)||Bρ(z)Uj|<β|Bρ|\alpha\,|B_{\rho/2}|\leq|U_{k}\cap B_{\rho/2}(w)|\leq|B_{\rho}(z)\setminus U_{j}|<\beta\,|B_{\rho}|

which gives a contradiction if β<α/2n+1\beta<\alpha/2^{n+1}. By Bρ/2(z)Uk=B_{\rho/2}(z)\cap\partial U_{k}=\varnothing it follows that

Bρ/2(z)cl(Uj)cl(E).B_{\rho/2}(z)\subset\mathrm{cl}\,(U_{j})\cup\mathrm{cl}\,(E)\,. (7.44)

Let us now set

E=EBρ/2(z),K=(KBρ/2(z))(E(1)Bρ/2(z)).E^{\prime}=E\setminus B_{\rho/2}(z)\,,\qquad K^{\prime}=\big{(}K\setminus B_{\rho/2}(z)\big{)}\cup\big{(}E^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho/2}(z)\big{)}\,. (7.45)

By (1.41), if n(EBρ/2)=0\mathcal{H}^{n}(\partial^{*}E\cap\partial B_{\rho/2})=0, then (K,E)𝒦(K^{\prime},E^{\prime})\in\mathcal{K}, since (ΩBρ/2(z))EKBρ/2(z)K(\Omega\setminus B_{\rho/2}(z))\cap\partial^{*}E\subset K\setminus B_{\rho/2}(z)\subset K^{\prime} implies

ΩE=nΩ{((E)Bρ/2(z))(E(1)Bρ/2(z))}K.\Omega\cap\partial^{*}E^{\prime}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\Omega\cap\big{\{}\big{(}(\partial^{*}E)\setminus B_{\rho/2}(z)\big{)}\cup\big{(}E^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho/2}(z)\big{)}\big{\}}\subset K^{\prime}\,.

Moreover K(E(1))K^{\prime}\cup(E^{\scriptscriptstyle{(1)}})^{\prime} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} since it contains (KE)Bρ/2(z)(K\cup E)\setminus B_{\rho/2}(z), and

(KE)Bρ/2(z) is 𝒞-spanning 𝐖.\mbox{$(K\cup E)\setminus B_{\rho/2}(z)$ is $\mathcal{C}$-spanning $\mathbf{W}$}\,. (7.46)

Indeed, if γ𝒞\gamma\in\mathcal{C} and γ(SS1)(KE)Bρ/2(z)=\gamma(\SS^{1})\cap(K\cup E)\setminus B_{\rho/2}(z)=\emptyset, then by applying Lemma 7.2 to S=KES=K\cup E and B=B2rB=B_{2\,r} we see that either γ(SS1)(KE)B2r\gamma(\SS^{1})\cap(K\cup E)\setminus B_{2\,r}\neq\varnothing (and thus γ(SS1)(KE)Bρ/2(z)\gamma(\SS^{1})\cap(K\cup E)\setminus B_{\rho/2}(z)\neq\varnothing by Bρ/2(z)BrB_{\rho/2}(z)\subset B_{r}), or there are khk\neq h such that γ(SS1)Uk\gamma(\SS^{1})\cap\partial U_{k}\neq\varnothing and γ(SS1)Uh\gamma(\SS^{1})\cap\partial U_{h}\neq\varnothing. Up to possibly switch kk and hh, we have that kjk\neq j, so that (7.44) implies that γ(SS1)Uk=γ(SS1)UkBρ/2(z)\varnothing\neq\gamma(\SS^{1})\cap\partial U_{k}=\gamma(\SS^{1})\cap\partial U_{k}\setminus B_{\rho/2}(z), where the latter set is contained in KBρ/2(z)K\setminus B_{\rho/2}(z) by (7.22) and Bρ/2(z)BrB_{\rho/2}(z)\subset B_{r}. This proves (7.46).

We can thus plug the competitor (K,E)(K^{\prime},E^{\prime}) defined in (7.45) into (7.39), and find

bk(K,E;Bρ/2(z))bk(K,E;cl(Bρ/2(z)))+Λ|EBρ/2(z)|,\mathcal{F}_{\rm bk}(K,E;B_{\rho/2}(z))\leq\mathcal{F}_{\rm bk}\big{(}K^{\prime},E^{\prime};\mathrm{cl}\,(B_{\rho/2}(z))\big{)}+\Lambda\,|E\cap B_{\rho/2}(z)|\,,

for every ρ<min{r0,dist(z,B2r)}\rho<\min\{r_{0},{\rm dist}(z,\partial B_{2\,r})\} such that n(KBρ/2(z))=0\mathcal{H}^{n}(K\cap\partial B_{\rho/2}(z))=0. Now, by Lemma 7.1 and by (7.44) we have

bk(K,E;Bρ/2(z))P(Uj;Bρ/2(z))=P(E;Bρ/2(z)),\mathcal{F}_{\rm bk}(K,E;B_{\rho/2}(z))\geq P(U_{j};B_{\rho/2}(z))=P(E;B_{\rho/2}(z))\,,

while (1.40) gives

cl(Bρ/2/z)K=ncl(Bρ/2/z)E=nE(1)Bρ/2(z),\mathrm{cl}\,(B_{\rho/2}/z)\cap K^{\prime}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}\mathrm{cl}\,(B_{\rho/2}/z)\cap\partial^{*}E^{\prime}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}E^{\scriptscriptstyle{(1)}}\cap\partial B_{\rho/2}(z)\,,

thus proving that, for a.e. ρ<min{r0,dist(z,B2r)}\rho<\min\{r_{0},{\rm dist}(z,\partial B_{2\,r})\},

P(E;Bρ/2(z))n(E(1)Bρ/2(z))+Λ|EBρ/2(z)|.P(E;B_{\rho/2}(z))\leq\mathcal{H}^{n}(E^{\scriptscriptstyle{(1)}}\cap B_{\rho/2}(z))+\Lambda\,|E\cap B_{\rho/2}(z)|\,.

Since zB2rUjz\in B_{2\,r}\cap\partial U_{j} and Bρ/2(z)Uj=Bρ/2(z)EB_{\rho/2}(z)\cap\partial^{*}U_{j}=B_{\rho/2}(z)\cap\partial^{*}E, by (1.35) we see that |EBρ/2(z)|>0|E\cap B_{\rho/2}(z)|>0 for every ρ<min{r0,dist(z,B2r)}\rho<\min\{r_{0},{\rm dist}(z,\partial B_{2\,r})\}. By a standard argument, up to further decrease the value of r0r_{0}, we find that for some α=α(n)\alpha^{\prime}=\alpha^{\prime}(n) it holds

|EBρ/2(z)|α|Bρ/2|,ρ<min{r0,dist(z,B2r)},|E\cap B_{\rho/2}(z)|\geq\alpha^{\prime}\,|B_{\rho/2}|\,,\qquad\forall\rho<\min\{r_{0},{\rm dist}(z,\partial B_{2\,r})\}\,,

and since |EBρ/2(z)|=|Bρ/2(z)Uj||E\cap B_{\rho/2}(z)|=|B_{\rho/2}(z)\setminus U_{j}| this give a contradiction with (7.43) up to further decrease the value of β\beta.

Step three: We prove (7.22) and (7.23). The lower bound in (7.32) implies (7.23), i.e., J=#{j:UjBr}<J=\#\{j:U_{j}\cap B_{r}\neq\varnothing\}<\infty. Next, by B2rUjKB_{2\,r}\cap\partial U_{j}\subset K (last inclusion in (7.27)), to prove (7.22) it suffices to show that

KBrj=1JUj.K\cap B_{r}\subset\cup_{j=1}^{J}\partial U_{j}\,. (7.47)

Now, if zKBrz\in K\cap B_{r}, then by KE=K\cap E=\varnothing we have either zKcl(E)z\in K\setminus\mathrm{cl}\,(E) or zBrEz\in B_{r}\cap\partial E, and, in the latter case, |EBρ(z)|(1c)|Bρ||E\cap B_{\rho}(z)|\leq(1-c)\,|B_{\rho}| for every ρ<min{r0,dist(z,𝐖)}\rho<\min\{r_{0},{\rm dist}(z,\partial\mathbf{W})\} thanks to (6.2). Therefore, in both cases, zz is an accumulation point for (j=1JUj)(1)Br(\cup_{j=1}^{J}U_{j})^{\scriptscriptstyle{(1)}}\cap B_{r}. Since JJ is finite, there must be at least one jj such that zcl(Uj)z\in\mathrm{cl}\,(U_{j}) – hence zUjz\in\partial U_{j} thanks to KUj=K\cap U_{j}=\varnothing.

Before moving to the next step, we also notice that

bk(K,E;Br)=j=1JP(Uj;Br).\mathcal{F}_{\rm bk}(K,E;B_{r})=\sum_{j=1}^{J}P(U_{j};B_{r})\,. (7.48)

Indeed, by (7.22), (7.23), and (7.31) we have

KBr=Brj=1JUj=nBrj=1JUj,K\cap B_{r}=B_{r}\cap\cup_{j=1}^{J}\partial U_{j}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}B_{r}\cap\cup_{j=1}^{J}\partial^{*}U_{j}\,, (7.49)

so that, in the application of Lemma 7.1, i.e. in (7.40), the multiplicity 22 terms vanishes, and we find (7.48).

Step four: In this step we consider a set of finite perimeter V1V_{1} such that, for some B:=Bρ(z)BrB:=B_{\rho}(z)\subset B_{r} with ρ<r0\rho<r_{0} and n(KB)=0\mathcal{H}^{n}(K\cap\partial B)=0, we have

U1ΔV1B.\displaystyle U_{1}\Delta V_{1}\subset\!\subset B\,. (7.50)

We then define a pair of Borel sets (K,E)(K^{\prime},E^{\prime}) as

E\displaystyle E^{\prime} =\displaystyle= (EB)[B(V1Δ(EU1))],\displaystyle\big{(}E\setminus B\big{)}\,\cup\,\big{[}B\cap\big{(}V_{1}\Delta(E\cup U_{1})\big{)}\big{]}\,, (7.51)
K\displaystyle K^{\prime} =\displaystyle= (KB)[B(V1U2UJ)],\displaystyle\big{(}K\setminus B\big{)}\,\cup\,\big{[}B\cap\big{(}\partial^{*}V_{1}\cup\partial^{*}U_{2}\cup\cdots\cup\partial^{*}U_{J}\big{)}\big{]}\,, (7.52)

and show that (K,E)𝒦B(K^{\prime},E^{\prime})\in\mathcal{K}_{\rm B}, K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, and

bk(K,E)bk(K,E)P(V1;B)P(U1;B).\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime})-\mathcal{F}_{\rm bk}(K,E)\leq P(V_{1};B)-P(U_{1};B)\,. (7.53)

As a consequence of (7.53), (7.33) and |EΔE|=|U1ΔV1||E\Delta E^{\prime}|=|U_{1}\Delta V_{1}|, we find of course that P(U1;Ω)P(V1;Ω)+Λ|U1ΔV1|P(U_{1};\Omega)\leq P(V_{1};\Omega)+\Lambda\,|U_{1}\Delta V_{1}|, thus showing that U1U_{1} is a (Λ,r0)(\Lambda,r_{0})-perimeter minimizer in Ω\Omega.

Proving that (K,E)𝒦B(K^{\prime},E^{\prime})\in\mathcal{K}_{\rm B} is immediately reduced to showing that BEB\cap\partial^{*}E^{\prime} is n\mathcal{H}^{n}-contained in B(V1U2UJ)B\cap(\partial^{*}V_{1}\cup\partial^{*}U_{2}\cup\cdots\cup\partial^{*}U_{J}) thanks to n(KB)=0\mathcal{H}^{n}(K\cap\partial B)=0. Now, on taking into account that, by (1.39) and (1.41), (XY)\partial^{*}(X\cup Y) and (XY)\partial^{*}(X\setminus Y) are both n\mathcal{H}^{n}-contained in (X)(Y)(\partial^{*}X)\cup(\partial^{*}Y), and thus (XΔY)\partial^{*}(X\Delta Y) is too, we easily see that

BE=B[V1Δ(EU1)]n(BV1)(B(EU1)).B\cap\partial^{*}E^{\prime}=B\cap\partial^{*}[V_{1}\Delta(E\cup U_{1})]\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}(B\cap\partial^{*}V_{1})\cup(B\cap\partial^{*}(E\cup U_{1}))\,.

However, B(EU1)=B(j=2JUj)B\cap(E\cup U_{1})=B\setminus(\cup_{j=2}^{J}U_{j}), so that X=(n+1X)\partial^{*}X=\partial^{*}(\mathbb{R}^{n+1}\setminus X) gives

B(EU1)=B(j=2JUj)nBj2Uj,B\cap\partial^{*}(E\cup U_{1})=B\cap\partial^{*}(\cup_{j=2}^{J}U_{j})\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}B\cap\cup_{j\geq 2}\partial^{*}U_{j}\,,

where we have used again the n\mathcal{H}^{n}-containment of (XY)\partial^{*}(X\cup Y) in (X)(Y)(\partial^{*}X)\cup(\partial^{*}Y). This proves that (K,E)𝒦B(K^{\prime},E^{\prime})\in\mathcal{K}_{\rm B}.

To prove that K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, we show that the set SS defined by

S=((KE)B)(cl(B)j2Uj),S=\big{(}(K\cup E)\setminus B\big{)}\cup\big{(}\mathrm{cl}\,(B)\cap\cup_{j\geq 2}\partial U_{j}\big{)}\,,

is n\mathcal{H}^{n}-contained in K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}} and is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}.

To prove that SS is n\mathcal{H}^{n}-contained in K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}}, we start by noticing that (KE)cl(B)(K\cup E)\setminus\mathrm{cl}\,(B) is n\mathcal{H}^{n}-equivalent to (KE(1)E)cl(B)KE(1)(K\cup E^{\scriptscriptstyle{(1)}}\cup\partial^{*}E)\setminus\mathrm{cl}\,(B)\subset K\cup E^{\scriptscriptstyle{(1)}} (by (K,E)𝒦B(K,E)\in\mathcal{K}_{\rm B}), whereas |(EΔE)B|=0|(E\Delta E^{\prime})\setminus B|=0 implies (E(1)Δ(E)(1))cl(B)=(E^{\scriptscriptstyle{(1)}}\Delta(E^{\prime})^{\scriptscriptstyle{(1)}})\setminus\mathrm{cl}\,(B)=\varnothing: hence Scl(B)S\setminus\mathrm{cl}\,(B) if n\mathcal{H}^{n}-contained in K(E)(1)K^{\prime}\cup(E^{\prime})^{\scriptscriptstyle{(1)}}. Next, by (7.31) and by definition of KK^{\prime},

SB=Bj2Uj=nBj2UjK.S\cap B=B\cap\cup_{j\geq 2}\partial U_{j}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}B\cap\cup_{j\geq 2}\partial^{*}U_{j}\subset K^{\prime}\,.

Finally, by n(KB)=0\mathcal{H}^{n}(K\cap\partial B)=0, (7.26), and Federer’s theorem, (SB)K(S\cap\partial B)\setminus K is n\mathcal{H}^{n}-equivalent to (E(1)B)K(E^{\scriptscriptstyle{(1)}}\cap\partial B)\setminus K, where E(1)A=(E)(1)AE^{\scriptscriptstyle{(1)}}\cap A=(E^{\prime})^{\scriptscriptstyle{(1)}}\cap A in an open neighborhood AA of B\partial B thanks to U1ΔV1BU_{1}\Delta V_{1}\subset\!\subset B.

To prove that SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, since SS is relatively closed in Ω\Omega and thanks to Theorem A.1, we only need to check that Sγ(SS1)S\cap\gamma(\SS^{1})\neq\varnothing for every γ𝒞\gamma\in\mathcal{C}. Since (KE)γ(SS1)(K\cup E)\cap\gamma(\SS^{1})\neq\varnothing for every γ𝒞\gamma\in\mathcal{C}, this is immediate unless γ\gamma is such that Sγ(SS1)B=S\cap\gamma(\SS^{1})\setminus B=\varnothing; in that case, however, Lemma 7.2 implies the existence of jkj\neq k such that γ(SS1)BUj\gamma(\SS^{1})\cap B\cap\partial U_{j} and γ(SS1)BUk\gamma(\SS^{1})\cap B\cap\partial U_{k} are both non-empty. Since either j2j\geq 2 or k2k\geq 2, we conclude by (7.26) that γ(SS1)BK\gamma(\SS^{1})\cap B\cap K^{\prime}\neq\varnothing, thus completing the proof.

We are thus left to prove the validity of (7.53). Keeping (7.48) and bk(K,E;B)bd(K,E;B)\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime};B)\leq\mathcal{F}_{\rm bd}(K^{\prime},E^{\prime};B) into account, this amounts to showing that

bd(K,E;B)=n(BE)+2n(BKE)=P(V1;B)+j=2JP(Uj;B).\mathcal{F}_{\rm bd}(K^{\prime},E^{\prime};B)=\mathcal{H}^{n}(B\cap\partial^{*}E^{\prime})+2\,\mathcal{H}^{n}\big{(}B\cap K^{\prime}\setminus\partial^{*}E^{\prime}\big{)}=P(V_{1};B)+\sum_{j=2}^{J}P(U_{j};B)\,. (7.54)

To this end we notice that by (1.44) and BE=B[V1Δ(EU1)]B\cap E^{\prime}=B\cap[V_{1}\Delta(E\cup U_{1})] we have

BE\displaystyle B\cap\partial^{*}E^{\prime} =n\displaystyle\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=} B{V1(EU1)}\displaystyle B\cap\big{\{}\partial^{*}V_{1}\cup\partial^{*}(E\cup U_{1})\big{\}}
=n\displaystyle\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=} B{(V1)(U1(0)E)(E(0)U1)},\displaystyle B\cap\big{\{}(\partial^{*}V_{1})\,\cup\,(U_{1}^{\scriptscriptstyle{(0)}}\cap\partial^{*}E)\,\cup\,(E^{\scriptscriptstyle{(0)}}\cap\partial^{*}U_{1})\big{\}}\,,

where we have used (1.39) and n({νE=νU1})=0\mathcal{H}^{n}(\{\nu_{E}=\nu_{U_{1}}\})=0 (as EU1=E\cap U_{1}=\varnothing). By (1.46) and (1.47), since {BE,BUj}j=1N\{B\cap E,B\cap U_{j}\}_{j=1}^{N} is a Caccioppoli partition of BB, we have

U1(0)E=(E)j2(Uj),E(0)U1=(U1)j2(Uj),U_{1}^{\scriptscriptstyle{(0)}}\cap\partial^{*}E=(\partial^{*}E)\cap\bigcup_{j\geq 2}(\partial^{*}U_{j})\,,\qquad E^{\scriptscriptstyle{(0)}}\cap\partial^{*}U_{1}=(\partial^{*}U_{1})\cap\bigcup_{j\geq 2}(\partial^{*}U_{j})\,,

so that

BE\displaystyle B\cap\partial^{*}E^{\prime} =n\displaystyle\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=} B{(V1)([(E)(U1)]j2(Uj))},\displaystyle B\cap\Big{\{}(\partial^{*}V_{1})\cup\Big{(}\big{[}(\partial^{*}E)\cup(\partial^{*}U_{1})\big{]}\cap\bigcup_{j\geq 2}(\partial^{*}U_{j})\Big{)}\Big{\}}\,,
B(KE)\displaystyle B\cap(K^{\prime}\setminus\partial^{*}E^{\prime}) =n\displaystyle\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=} B(j2Uj)[(E)(U1)].\displaystyle B\cap\Big{(}\bigcup_{j\geq 2}\partial^{*}U_{j}\Big{)}\setminus\big{[}(\partial^{*}E)\cup(\partial^{*}U_{1})\big{]}\,.

We thus find

n(BE)+2n(B(KE))\displaystyle\mathcal{H}^{n}(B\cap\partial^{*}E)+2\,\mathcal{H}^{n}(B\cap(K^{\prime}\setminus\partial^{*}E^{\prime}))
=P(V1;B)+2n((j2Uj)(EU1))+n((j2Uj)(EU1))\displaystyle=P(V_{1};B)+2\,\mathcal{H}^{n}\Big{(}\Big{(}\bigcup_{j\geq 2}\partial^{*}U_{j}\Big{)}\setminus(\partial^{*}E\cup\partial^{*}U_{1})\Big{)}+\mathcal{H}^{n}\Big{(}\Big{(}\bigcup_{j\geq 2}\partial^{*}U_{j}\Big{)}\cap(\partial^{*}E\cup\partial^{*}U_{1})\Big{)}
=P(V1;B)+j2P(Uj;B),\displaystyle=P(V_{1};B)+\sum_{j\geq 2}P(U_{j};B)\,,

that is (7.54).

Step five: In this final step we prove conclusions (iv) and (v). To this end we fix x[Ω(EE)]Σx\in[\Omega\cap(\partial E\setminus\partial^{*}E)]\setminus\Sigma, and recall that, by conclusion (iv)α, there are r>0r>0, νSSn\nu\in\SS^{n}, u1,u2C1,α(𝐃rν(x);(r/4,r/4))u_{1},u_{2}\in C^{1,\alpha}(\mathbf{D}_{r}^{\nu}(x);(-r/4,r/4)) (α(0,1/2)\alpha\in(0,1/2) arbitrary) such that u1(x)=u2(x)=0u_{1}(x)=u_{2}(x)=0, u1u2u_{1}\leq u_{2} on 𝐃rν(x)\mathbf{D}_{r}^{\nu}(x), {u1<u2}\{u_{1}<u_{2}\} and int{u1=u2}{\rm int}\{u_{1}=u_{2}\} are both non-empty, and

𝐂rν(x)K\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap K =\displaystyle= i=1,2{y+ui(y)ν:y𝐃rν(x)},\displaystyle\cup_{i=1,2}\big{\{}y+u_{i}(y)\,\nu:y\in\mathbf{D}_{r}^{\nu}(x)\big{\}}\,, (7.55)
𝐂rν(x)E\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap\partial^{*}E =\displaystyle= i=1,2{y+ui(y)ν:y{u1<u2}},\displaystyle\cup_{i=1,2}\big{\{}y+u_{i}(y)\nu:y\in\{u_{1}<u_{2}\}\big{\}}\,, (7.56)
𝐂rν(x)E\displaystyle\mathbf{C}_{r}^{\nu}(x)\cap E =\displaystyle= {y+tν:y{u1<u2},u1(x)<t<u2(x)}.\displaystyle\big{\{}y+t\,\nu:y\in\{u_{1}<u_{2}\}\,,u_{1}(x)<t<u_{2}(x)\big{\}}\,. (7.57)

We claim that (u1,u2)(u_{1},u_{2}) has the minimality property

𝒜(u1,u2)𝒜(w1,w2):=𝐃rν(x)1+|w1|2+1+|w2|2,\mathcal{A}(u_{1},u_{2})\leq\mathcal{A}(w_{1},w_{2}):=\int_{\mathbf{D}_{r}^{\nu}(x)}\sqrt{1+|\nabla w_{1}|^{2}}+\sqrt{1+|\nabla w_{2}|^{2}}\,, (7.58)

among all pairs (w1,w2)(w_{1},w_{2}) with w1,w2Lip(𝐃rν(x);(r/2,r/2))w_{1},w_{2}\in{\rm Lip}(\mathbf{D}_{r}^{\nu}(x);(-r/2,r/2)) that satisfy

{w1w2,on 𝐃rν(x),wk=uk,on 𝐃rν(x)k=1,2,𝐃rν(x)w2w1=𝐃rν(x)u2u1.\left\{\begin{split}&w_{1}\leq w_{2}\,,\quad\mbox{on $\mathbf{D}_{r}^{\nu}(x)$}\,,\\ &w_{k}=u_{k}\,,\quad\mbox{on $\partial\mathbf{D}_{r}^{\nu}(x)$, $k=1,2$}\,,\end{split}\right.\qquad\int_{\mathbf{D}_{r}^{\nu}(x)}w_{2}-w_{1}=\int_{\mathbf{D}_{r}^{\nu}(x)}u_{2}-u_{1}\,. (7.59)

Indeed, starting from a given a pair (w1,w2)(w_{1},w_{2}) as in (7.59), we can define (K𝐂rν(x),E𝐂rν(x))(K^{\prime}\cap\mathbf{C}_{r}^{\nu}(x),E^{\prime}\cap\mathbf{C}_{r}^{\nu}(x)) by replacing (u1,u2)(u_{1},u_{2}) with (w1,w2)(w_{1},w_{2}) in (7.55) and (7.57), and then define (K,E)𝒦B(K^{\prime},E^{\prime})\in\mathcal{K}_{\rm B} by setting K𝐂rν(x)=K𝐂rν(x)K^{\prime}\setminus\mathbf{C}_{r}^{\nu}(x)=K\setminus\mathbf{C}_{r}^{\nu}(x) and E𝐂rν(x)=E𝐂rν(x)E^{\prime}\setminus\mathbf{C}_{r}^{\nu}(x)=E\setminus\mathbf{C}_{r}^{\nu}(x). Since 𝐂rν(KE)=𝐂rν(KE)\partial\mathbf{C}_{r}^{\nu}\setminus(K^{\prime}\cup E^{\prime})=\partial\mathbf{C}_{r}^{\nu}\setminus(K\cup E) it is easily seen (by a simple modification of Lemma 7.2 where balls are replaced by cylinders) that (K,E)(K^{\prime},E^{\prime}) is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}. Since |E|=|E||E^{\prime}|=|E|, the minimality of (K,E)(K,E) in Ψbk(v)\Psi_{\rm bk}(v) implies that bk(K,E)bk(K,E)\mathcal{F}_{\rm bk}(K,E)\leq\mathcal{F}_{\rm bk}(K^{\prime},E^{\prime}), which readily translates into (7.58).

Recalling that both A0=int{u1=u2}A_{0}={\rm int}\{u_{1}=u_{2}\} and A+={u1<u2}A_{+}=\{u_{1}<u_{2}\} are non-empty open subsets of 𝐃rν(x)\mathbf{D}_{r}^{\nu}(x), and denoting by MS(u)[φ]=𝐃rν(x)φ[(u)/1+|u|2]{\rm MS}(u)[\varphi]=\int_{\mathbf{D}_{r}^{\nu}(x)}\nabla\varphi\cdot[(\nabla u)/\sqrt{1+|\nabla u|^{2}}] the distributional mean curvature operator, we find that

MS(u1)+MS(u2)=0,\displaystyle{\rm MS}(u_{1})+{\rm MS}(u_{2})=0\,, on 𝐃rν(x),\displaystyle\qquad\mbox{on $\mathbf{D}_{r}^{\nu}(x)$}\,,
MS(uk)=0,\displaystyle{\rm MS}(u_{k})=0\,, on A0 for each k=1,2,\displaystyle\qquad\mbox{on $A_{0}$ for each $k=1,2$}\,,
MS(u2)=MS(u1)=λ,\displaystyle{\rm MS}(u_{2})=-{\rm MS}(u_{1})=\lambda\,, on A+,\displaystyle\qquad\mbox{on $A_{+}$}\,, (7.60)

for some constant λ\lambda\in\mathbb{R}; in particular, u1,u2C(A0)C(A+)u_{1},u_{2}\in C^{\infty}(A_{0})\cap C^{\infty}(A_{+}). We notice that it must be

λ<0.\lambda<0\,. (7.61)

Indeed, arguing by contradiction, should it be that λ0\lambda\geq 0, then by (7.60) we find MS(u2)0{\rm MS}(u_{2})\geq 0 and MS(u1)0{\rm MS}(u_{1})\leq 0 on A+A_{+}. Since A+A_{+} is open an non-empty, there is an open ball BA+B\subset A_{+} such that BA+={y0}\partial B\cap\partial A_{+}=\{y_{0}\}. Denoting by x0x_{0} the center of BB and setting ν=(x0y0)/|x0y0|\nu=(x_{0}-y_{0})/|x_{0}-y_{0}|, by u1u2u_{1}\leq u_{2}, u1(y0)=u2(y0)u_{1}(y_{0})=u_{2}(y_{0}) and ukC1(𝐃rν(x))u_{k}\in C^{1}(\mathbf{D}_{r}^{\nu}(x)) we find that u1(y0)=u2(y0)\nabla u_{1}(y_{0})=\nabla u_{2}(y_{0}). At the same time, by applying Hopf’s lemma in BB at y0y_{0}, we see that since MS(u2)0{\rm MS}(u_{2})\geq 0 and MS(u1)0{\rm MS}(u_{1})\leq 0 on BB, it must be νu2(y0)<0\nu\cdot\nabla u_{2}(y_{0})<0 and νu1(y0)>0\nu\cdot\nabla u_{1}(y_{0})>0, against u1(y0)=u2(y0)\nabla u_{1}(y_{0})=\nabla u_{2}(y_{0}).

By (7.60), (7.61), and u2u1u_{2}\geq u_{1} on 𝐃rν(x)\mathbf{D}_{r}^{\nu}(x) we can apply the sharp regularity theory for the double membrane problem developed in [Sil05, Theorem 5.1] and deduce that u1,u2C1,1(𝐃rν(x))u_{1},u_{2}\in C^{1,1}(\mathbf{D}_{r}^{\nu}(x)). Next we notice that, for every φCc(A+)\varphi\in C^{\infty}_{c}(A_{+}), and setting u+=u2u1u_{+}=u_{2}-u_{1},

2λA+φ=MS(u2)[φ]MS(u1)[φ]=A+A(x)[u+]φ,2\,\lambda\,\int_{A_{+}}\varphi={\rm MS}(u_{2})[\varphi]-{\rm MS}(u_{1})[\varphi]=\int_{A_{+}}{\rm A}(x)[\nabla u_{+}]\cdot\nabla\varphi\,,

where we have set, with f(z)=1+|z|2f(z)=\sqrt{1+|z|^{2}},

A(x)=012f(su2(x)+(1s)u1(x))𝑑s.{\rm A}(x)=\int_{0}^{1}\,\nabla^{2}f\big{(}s\,\nabla u_{2}(x)+(1-s)\,\nabla u_{1}(x)\big{)}\,ds\,.

In particular, u+C1,1(𝐃rν(x))u_{+}\in C^{1,1}(\mathbf{D}_{r}^{\nu}(x)) is a non-negative distributional solution of

div(A(x)u+)=2λ,on A+,{\rm div}\,({\rm A}(x)\nabla u_{+})=-2\,\lambda\,,\qquad\mbox{on $A_{+}$}\,,

with a strictly positive right-hand side (by (7.61)) and with ALip(A+;symn×n){\rm A}\in{\rm Lip}(A_{+};\mathbb{R}^{n\times n}_{\rm sym}) uniformly elliptic. We can thus apply the regularity theory for free boundaries developed in [FGS15, Theorem 1.1, Theorem 4.14] to deduce that

FB=𝐃rν(x){u+=0}=𝐃rν(x){u2=u1},{\rm FB}=\mathbf{D}_{r}^{\nu}(x)\cap\partial\{u_{+}=0\}=\mathbf{D}_{r}^{\nu}(x)\cap\partial\{u_{2}=u_{1}\}\,,

can be partitioned into sets Reg{\rm Reg} and Sing{\rm Sing} such that Reg{\rm Reg} is relatively open in FB{\rm FB} and such that for every zRegz\in{\rm Reg} there are r>0r>0 and β(0,1)\beta\in(0,1) such that Br(x)FBB_{r}(x)\cap{\rm FB} is a C1,βC^{1,\beta}-embedded (n1)(n-1)-dimensional manifold, and such that Sing=k=0n1Singk{\rm Sing}=\cup_{k=0}^{n-1}{\rm Sing}_{k} is relatively closed in FB{\rm FB}, with each Singk{\rm Sing}_{k} locally k\mathcal{H}^{k}-rectifiable in 𝐃rν(x)\mathbf{D}_{r}^{\nu}(x). Since, by (7.56),

𝐂rν(x)(EE)={y+u1(y)ν:yFB}\mathbf{C}_{r}^{\nu}(x)\cap(\partial E\setminus\partial^{*}E)=\big{\{}y+u_{1}(y)\,\nu:y\in{\rm FB}\big{\}}

and u1C1,1(𝐃rν(x))u_{1}\in C^{1,1}(\mathbf{D}_{r}^{\nu}(x)), we conclude by a covering argument that Ω(EE)\Omega\cap(\partial E\setminus\partial^{*}E) has all the required properties, and complete the proof of the theorem. ∎

8. Equilibrium across transition lines in wet foams (Theorem 1.7)

Proof of Theorem 1.7.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1} be open and let (K,E)𝒦foam(K_{*},E_{*})\in\mathcal{K}_{\rm foam}. We can find (K,E)𝒦(K,E)\in\mathcal{K} such that KK is n\mathcal{H}^{n}-equivalent to KK_{*}, EE Lebesgue equivalent to EE_{*}, and KE(1)=K\cap E^{\scriptscriptstyle{(1)}}=\varnothing by repeating with minor variations the considerations made in step one of the proof of Theorem 6.2 (we do not have to worry about the 𝒞\mathcal{C}-spanning condition, but have to keep track of the volume constraint imposed for each UiU_{i}, which can be done by using the volume-fixing variations for clusters from [Mag12, Part IV]). In proving the regularity part of the statement, thanks to Theorem 2.1-(a) we can directly work with balls BΩB\subset\!\subset\Omega having radius less than r0r_{0} (with r0r_{0} as in (1.33)), and consider the open connected components {Ui}i\{U_{i}\}_{i} of BB induced by KEK\cup E. Using Lemma 7.1 and, again, volume-fixing variation techniques in place of the theory of homotopic spanning, we can proceed to prove analogous statement to (7.8), (7.9), (7.10), and (7.11), thus proving the (Λ,r0)(\Lambda,r_{0})-minimality of each UiU_{i} in BB. The claimed C1,αC^{1,\alpha}-regularity of each UiU_{i} outside of a closed set Σ\Sigma with the claimed dimensional estimates follows then from De Giorgi’s theory of perimeter minimality [DG60, Tam82, Mag12]. ∎

Appendix A Equivalence of homotopic spanning conditions

In Theorem A.1 we prove that, when SS is a closed set, the notion of “SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}” introduced in Definition B boils down to the one in Definition A. We then show that the property of being 𝒞\mathcal{C}-spanning is stable under reduction to the rectifiable part of a Borel set, see Lemma 2.2.

Theorem A.1.

Given a closed set 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1}, a spanning class 𝒞\mathcal{C} for 𝐖\mathbf{W}, and a set SS relatively closed in Ω\Omega, the following two properties are equivalent:

(i): for every γ𝒞\gamma\in\mathcal{C}, we have Sγ(SS1)S\cap\gamma(\SS^{1})\neq\varnothing;

(ii): for every (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}) and for 1\mathcal{H}^{1}-a.e. s𝕊1s\in\mathbb{S}^{1}, we have

for n-a.e. xT[s],\displaystyle\mbox{for $\mathcal{H}^{n}$-a.e. $x\in T[s]$}\,, (A.1)
 a partition {T1,T2} of T with xeT1eT2,\displaystyle\mbox{$\exists$ a partition $\{T_{1},T_{2}\}$ of $T$ with $x\in\partial^{e}T_{1}\cap\partial^{e}T_{2}$}\,,
and s.t. ST[s] essentially disconnects T into {T1,T2}.\displaystyle\mbox{and s.t. $S\cup T[s]$ essentially disconnects $T$ into $\{T_{1},T_{2}\}$}\,.

In particular, SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} according to Definition A if and only if it does so according to Definition B.

Remark A.2 (xx-dependency of {T1,T2}\{T_{1},T_{2}\}).

In the situation of Figure 1.4 it is clear that the same choice of {T1,T2}\{T_{1},T_{2}\} can be used to check the validity of (A.1) at every xT[s]x\in T[s]. One may thus wonder if it could suffice to reformulate (A.1) so that the partition {T1,T2}\{T_{1},T_{2}\} is independent of xx. The simpler example we are aware of and that shows this simpler definition would not work is as follows. In 3\mathbb{R}^{3}, let 𝐖\mathbf{W} be a closed δ\delta-neighborhood of a circle Γ\Gamma, let UU be the open δ\delta-neighborhood of a loop with link number three (or higher odd number) with respect to 𝐖\mathbf{W}, let KK be the disk spanned by Γ\Gamma, and let S=Ω[(KU)U]S=\Omega\cap[(K\setminus U)\cup\partial U], see

Refer to caption
D1D_{1}D3D_{3}D2D_{2}A3A_{3}A2A_{2}𝐖\mathbf{W}UT[s]U\setminus T[s]SSA1A_{1}
Figure A.1. The situation in Remark A.2. The components A1A_{1}, A2A_{2} and A3A_{3} (depicted in purple, yellow, and green respectively) of UT[s]U\setminus T[s] are bounded by the three disks {Di}i=13\{D_{i}\}_{i=1}^{3} (depicted as boldface segments).

Figure A.1. Now consider a “test tube” TT which compactly contains UU and is such that, for every ss, UT[s]U\cap T[s] consists of three disks {Di}i=13\{D_{i}\}_{i=1}^{3}. Since UTU\subset T, the property “ST[s]S\cup T[s] essentially disconnects TT into {T1,T2}\{T_{1},T_{2}\} in such a way that T[s]TeT1eT2T[s]\subset T\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}” would immediately imply “U(ST[s])=UT[s]U\cap(S\cup T[s])=U\cap T[s] essentially disconnects TU=UT\cap U=U into {U1,U2}\{U_{1},U_{2}\} with UT[s]UeU1eU2U\cap T[s]\subset U\cap\partial^{e}U_{1}\cap\partial^{e}U_{2}”, where Ui=TiUU_{i}=T_{i}\cap U (see step one in the proof of Theorem 3.1 for a formal proof of this intuitive assertion). However, the latter property does not hold. To see this, denoting by {Ai}i=13\{A_{i}\}_{i=1}^{3} the three connected components of UT[s]U\setminus T[s], we would have U1=AiAjU_{1}=A_{i}\cup A_{j} and U2=AkU_{2}=A_{k} for some choice of ijkii\neq j\neq k\neq i, whereas, independently of the choice made, UeU1eU2U\cap\partial^{e}U_{1}\cap\partial^{e}U_{2} always fails to contain one of the disks {Di}i=13\{D_{i}\}_{i=1}^{3}: for example, if U1=A1A2U_{1}=A_{1}\cup A_{2} and U2=A3U_{2}=A_{3}, then UeU1eU2=D2D3U\cap\partial^{e}U_{1}\cap\partial^{e}U_{2}=D_{2}\cup D_{3}, and D1D_{1} is entirely missed. We conclude that the set SS just constructed, although clearly 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} in terms of Definition A, fails to satisfy the variant of (A.1) where a same partition {T1,T2}\{T_{1},T_{2}\} is required to work for n\mathcal{H}^{n}-a.e. choice of xT[s]x\in T[s].

Proof of Theorem A.1.

Step one: We prove that (ii) implies (i). Indeed, if there is γ𝒞\gamma\in\mathcal{C} such that Sγ(SS1)=S\cap\gamma(\SS^{1})=\varnothing, then, SS being closed, we can find (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}) such that dist(S,T)>0{\rm dist}(S,T)>0. By (ii), there is sSS1s\in\SS^{1} such that ST[s]S\cup T[s] essentially disconnects TT. By dist(S,T)>0{\rm dist}(S,T)>0 we see that (ST[s])T=T[s](S\cup T[s])\cap T=T[s], so that T[s]T[s] essentially disconnects TT, a contradiction.

Step two: We now prove that (i) implies (ii). To this end we consider an arbitrary (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}) and aim at proving the existence of JJ of full 1\mathcal{H}^{1}-measure in SS1\SS^{1} such that, if sJs\in J, then (A.1) holds.

This is trivial, with J=SS1J=\SS^{1}, if |ST|=|T||S\cap T|=|T|. Indeed, in this case, we have T=S(1)TT=S^{{\scriptscriptstyle{(1)}}}\cap T, that, combined with SS being closed, implies T=STT=S\cap T. In particular, ST[s]=TS\cup T[s]=T for every sSS1s\in\SS^{1}, and since, trivially, TT essentially disconnects TT, the conclusion follows.

We thus assume that |ST|<|T||S\cap T|<|T|: in particular,

U=TSU=T\setminus S

is a non-empty, open set, whose connected components are denoted by {Ui}iI\{U_{i}\}_{i\in I} (II a countable set). By the Lebesgue points theorem, n+1\mathcal{L}^{n+1}-a.e. xTx\in T belongs either to U(0)U^{\scriptscriptstyle{(0)}} or to UU. Then, by the smoothness of Φ\Phi and by the area formula, we can find a set JJ of full 1\mathcal{H}^{1}-measure in SS1\SS^{1} such that

n(T[s](U(0)U))=0,sJ.\mathcal{H}^{n}\big{(}T[s]\setminus(U^{\scriptscriptstyle{(0)}}\cup U)\big{)}=0\,,\qquad\forall s\in J\,. (A.2)

In particular, given sJs\in J, we just need to prove (A.1) when either xT[s]U(0)x\in T[s]\cap U^{\scriptscriptstyle{(0)}} or xT[s]Ux\in T[s]\cap U. Before examining these two cases we also notice that we can further impose on JJ that

n(T[s][eUeS(U(1)U)iI(Ui(1)Ui)])=0,sJ.\displaystyle\mathcal{H}^{n}\Big{(}T[s]\cap\Big{[}\partial^{e}U\cup\partial^{e}S\cup\big{(}U^{{\scriptscriptstyle{(1)}}}\setminus U\big{)}\cup\bigcup_{i\in I}\big{(}U_{i}^{{\scriptscriptstyle{(1)}}}\setminus U_{i}\big{)}\Big{]}\Big{)}=0\,,\qquad\forall s\in J\,. (A.3)

Indeed, again by the Lebesgue points theorem, the sets eU\partial^{e}U, eS\partial^{e}S, U(1)UU^{{\scriptscriptstyle{(1)}}}\setminus U, and iIUi(1)Ui\cup_{i\in I}U_{i}^{{\scriptscriptstyle{(1)}}}\setminus U_{i} are all n+1\mathcal{L}^{n+1}-negligible.

Case one, xT[s]U(0)x\in T[s]\cap U^{\scriptscriptstyle{(0)}}: To fix ideas, notice that U(0)U^{\scriptscriptstyle{(0)}}\neq\varnothing implies |ST|>0|S\cap T|>0, and in particular SS has positive Lebesgue measure. Given an arbitrary sJ{s}s^{\prime}\in J\setminus\{s\} we denote by {I1,I2}\{I_{1},I_{2}\} the partition of SS1\SS^{1} bounded by {s,s}\{s,s^{\prime}\}, and then consider the Borel sets

T1=Φ(I1×B1n)S,T2=Φ(I2×B1n)(Φ(I1×B1n)S).T_{1}=\Phi(I_{1}\times B_{1}^{n})\cap S\,,\qquad T_{2}=\Phi(I_{2}\times B_{1}^{n})\cup\,\Big{(}\Phi(I_{1}\times B_{1}^{n})\setminus S\Big{)}\,.

We first notice that {T1,T2}\{T_{1},T_{2}\} is a non-trivial partition of TT: Indeed |T1|>0|T_{1}|>0 since xx has density 1/21/2 for Φ(I1×B1n)\Phi(I_{1}\times B_{1}^{n}) and (by xU(0)x\in U^{\scriptscriptstyle{(0)}}) density 11 for STS\cap T; at the same time |T2|=|TT1||TS|>0|T_{2}|=|T\setminus T_{1}|\geq|T\setminus S|>0. Next, we claim that

T(1)eT1eT2 is n-contained in S.\mbox{$T^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}$ is $\mathcal{H}^{n}$-contained in $S$}\,. (A.4)

Indeed, since Φ(I1×B1n)\Phi(I_{1}\times B_{1}^{n}) is an open subset of TT with T[Φ(I1×B1n)]=T[s]T[s]T\cap\partial[\Phi(I_{1}\times B_{1}^{n})]=T[s]\cup T[s^{\prime}], and since eT1\partial^{e}T_{1} coincides with eS\partial^{e}S inside the open set Φ(I1×B1n)\Phi(I_{1}\times B_{1}^{n}), we easily see that

T(1)eT1eT2\displaystyle T^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}T_{1}\cap\partial^{e}T_{2} =\displaystyle= TeT1=Te(Φ(I1×B1n)S)\displaystyle T\cap\partial^{e}T_{1}=T\cap\partial^{e}\big{(}\Phi(I_{1}\times B_{1}^{n})\cap S\big{)}
\displaystyle\subset (Φ(I1×B1n)eS)((T[s]T[s])S(0)).\displaystyle\big{(}\Phi(I_{1}\times B_{1}^{n})\cap\partial^{e}S\big{)}\cup\Big{(}\big{(}T[s]\cup T[s^{\prime}]\big{)}\setminus S^{\scriptscriptstyle{(0)}}\Big{)}\,.

Now, on the one hand, by n(eS(T[s]T[s]))=0\mathcal{H}^{n}(\partial^{e}S\cap(T[s]\cup T[s^{\prime}]))=0 (recall (A.3)), it holds

(T[s]T[s])S(0) is n-contained in TS(1);\mbox{$\big{(}T[s]\cup T[s^{\prime}]\big{)}\setminus S^{\scriptscriptstyle{(0)}}$ is $\mathcal{H}^{n}$-contained in $T\cap S^{{\scriptscriptstyle{(1)}}}$}\,;

while, on the other hand, by ΩeSΩSΩS\Omega\cap\partial^{e}S\subset\Omega\cap\partial S\subset\Omega\cap S (since SS is closed in Ω\Omega) and by Φ(I1×B1n)TΩ\Phi(I_{1}\times B_{1}^{n})\subset T\subset\Omega, we also have that Φ(I1×B1n)eSTS\Phi(I_{1}\times B_{1}^{n})\cap\partial^{e}S\subset T\cap S; therefore

T(1)eT1eT2 is n-contained in T(SS(1))=TS,\mbox{$T^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}$ is $\mathcal{H}^{n}$-contained in $T\cap(S\cup S^{{\scriptscriptstyle{(1)}}})=T\cap S$}\,,

where we have used that SS is closed to infer S(1)SS^{{\scriptscriptstyle{(1)}}}\subset S. Having proved (A.4) and the non-triviality of {T1,T2}\{T_{1},T_{2}\}, we conclude that SS (and, thus, ST[s]S\cup T[s]) essentially disconnects TT into {T1,T2}\{T_{1},T_{2}\}. We are left to prove that xTeT1eT2x\in T\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}. To this end, we notice that xT[s](TS)(0)x\in T[s]\cap(T\setminus S)^{\scriptscriptstyle{(0)}} and Φ(I1×B1n)T\Phi(I_{1}\times B_{1}^{n})\subset T imply

|T1Br(x)|=|Φ(I1×B1n)SBr(x)|=|Φ(I1×B1n)Br(x)|+o(rn+1)=|Br(x)|2+o(rn+1),|T_{1}\cap B_{r}(x)|=|\Phi(I_{1}\times B_{1}^{n})\cap S\cap B_{r}(x)|=|\Phi(I_{1}\times B_{1}^{n})\cap B_{r}(x)|+{\rm o}(r^{n+1})=\frac{|B_{r}(x)|}{2}+{\rm o}(r^{n+1})\,,

so that x(T1)(1/2)eT1x\in(T_{1})^{\scriptscriptstyle{(1/2)}}\subset\partial^{e}T_{1}; since TeT1=TeT1eT2T\cap\partial^{e}T_{1}=T\cap\partial^{e}T_{1}\cap\partial^{e}T_{2} and xTx\in T we conclude the proof in the case when xT[s]U(0)x\in T[s]\cap U^{\scriptscriptstyle{(0)}}.

Case two, xT[s]Ux\in T[s]\cap U: In this case there exists iIi\in I such that xUix\in U_{i}, and, correspondingly, we claim that

{V1,V2} a non-trivial Borel partition of UiT[s],\displaystyle\mbox{$\exists\{V_{1},V_{2}\}$ a non-trivial Borel partition of $U_{i}\setminus T[s]$}\,, (A.5)
s.t. xeV1eV2 and T(V1V2)ST[s].\displaystyle\mbox{s.t. $x\in\partial^{e}V_{1}\cap\partial^{e}V_{2}$ and $T\cap(\partial V_{1}\cup\partial V_{2})\subset S\cup T[s]$}\,.

Given the claim, we conclude by setting T1=V1T_{1}=V_{1} and T2=V2(TUi)T_{2}=V_{2}\cup(T\setminus U_{i}). Indeed, since V2Ui=T2UiV_{2}\cap U_{i}=T_{2}\cap U_{i} with UiU_{i} open implies UieV1=UieT1U_{i}\cap\partial^{e}V_{1}=U_{i}\cap\partial^{e}T_{1}, we deduce from (A.5) that

xUieV1eV2=UieT1eT2;x\in U_{i}\cap\partial^{e}V_{1}\cap\partial^{e}V_{2}=U_{i}\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}\,;

at the same time, ST[s]S\cup T[s] essentially disconnects TT into {T1,T2}\{T_{1},T_{2}\} since, again by (A.5),

T(1)eT1eT2=TeT1=TeV1TV1ST[s].T^{{\scriptscriptstyle{(1)}}}\cap\partial^{e}T_{1}\cap\partial^{e}T_{2}=T\cap\partial^{e}T_{1}=T\cap\partial^{e}V_{1}\subset T\cap\partial V_{1}\subset S\cup T[s]\,.

We are thus left to prove (A.5). To this end, let us choose r(x)>0r(x)>0 small enough to have that Br(x)(x)UiB_{r(x)}(x)\subset U_{i}, and that Br(x)(x)T[s]B_{r(x)}(x)\setminus T[s] consists of exactly two connected components {V1x,V2x}\{V_{1}^{x},V_{2}^{x}\}; in this way,

x(V1x)(1/2)(V2x)(1/2).x\in(V_{1}^{x})^{\scriptscriptstyle{(1/2)}}\cap(V_{2}^{x})^{\scriptscriptstyle{(1/2)}}\,. (A.6)

Next, we define

V1= the connected component of UiT[s] containing V1x,\displaystyle\mbox{$V_{1}=$ the connected component of $U_{i}\setminus T[s]$ containing $V_{1}^{x}$}\,,
V2=Ui(T[s]V1).\displaystyle V_{2}=U_{i}\setminus(T[s]\cup V_{1})\,.

Clearly {V1,V2}\{V_{1},V_{2}\} is a partition of UiT[s]U_{i}\setminus T[s], and, thanks to V1V2T[s]Ui\partial V_{1}\cup\partial V_{2}\subset T[s]\cup\partial U_{i}, we have

T(V1V2)T(T[s]Ui)ST[s].T\cap(\partial V_{1}\cup\partial V_{2})\subset T\cap(T[s]\cup\partial U_{i})\subset S\cup T[s]\,.

Therefore (A.5) follows by showing that |V1||V2|>0|V_{1}|\,|V_{2}|>0. Since V1V_{1} contains the connected component V1xV_{1}^{x} of Br(x)(x)T[s]B_{r(x)}(x)\setminus T[s], which is open and non-empty, we have |V1|>0|V_{1}|>0. Arguing by contradiction, we assume that

|V2|=|Ui(T[s]V1)|=0.|V_{2}|=|U_{i}\setminus(T[s]\cup V_{1})|=0\,.

Since V1V_{1} is a connected component of the open set UiT[s]U_{i}\setminus T[s] this implies that

UiT[s]=V1.U_{i}\setminus T[s]=V_{1}\,.

Let x1V1xx_{1}\in V_{1}^{x} and x2V2xx_{2}\in V_{2}^{x} (where V1xV_{1}^{x} and V2xV_{2}^{x} are the two connected components of Br(x)(x)T[s]B_{r(x)}(x)\setminus T[s]). Since V1V_{1} is connected and {x1,x2}UiT[s]=V1\{x_{1},x_{2}\}\subset U_{i}\setminus T[s]=V_{1}, there is a smooth embedding γ1\gamma_{1} of [0,1][0,1] into V1V_{1} with γ1(0)=x1\gamma_{1}(0)=x_{1} and γ1(1)=x2\gamma_{1}(1)=x_{2}. Arguing as in [DLGM17b, Proof of Lemma 10, Step 2] using Sard’s theorem, we may modify γ1\gamma_{1} by composing with a smooth diffeomorphism such that the modified γ1\gamma_{1} intersects Br(x)(x)\partial B_{r(x)}(x) transversally at finitely many points. Thus γ1([0,1])clBr(x)(x)\gamma_{1}([0,1])\setminus\mathrm{cl}\,B_{r(x)}(x) is partitioned into finitely many curves γ1((ai,bi))\gamma_{1}((a_{i},b_{i})) for disjoint arcs (ai,bi)[0,1](a_{i},b_{i})\subset[0,1]. Since Br(x)(x)T[s]B_{r(x)}(x)\setminus T[s] is disconnected into V1xV_{1}^{x} and V2xV_{2}^{x} and γ1\gamma_{1} is disjoint from T[s]T[s], there exists ii such that, up to interchanging V1xV_{1}^{x} and V2xV_{2}^{x}, γ(ai)clV1xBr(x)(x)\gamma(a_{i})\in\mathrm{cl}\,V_{1}^{x}\cap\partial B_{r(x)}(x) and γ(bi)clV2xBr(x)(x)\gamma(b_{i})\in\mathrm{cl}\,V_{2}^{x}\cap\partial B_{r(x)}(x). Let us call γ1~\tilde{\gamma_{1}} the restriction of γ1\gamma_{1} to [ai,bi][a_{i},b_{i}]. Next, we choose a smooth embedding γ2\gamma_{2} of [0,1][0,1] into Br(x)(x)B_{r(x)}(x) such that γ2(0)=γ~1(ai)\gamma_{2}(0)=\tilde{\gamma}_{1}(a_{i}), γ2(1)=γ~1(bi)\gamma_{2}(1)=\tilde{\gamma}_{1}(b_{i}), and γ2([0,1])\gamma_{2}([0,1]) intersects T[s]Br(x)(x)T[s]\cap B_{r(x)}(x) at exactly one point, denoted by x12=γ2(t0)x_{12}=\gamma_{2}(t_{0}), with

γ2(t0)0.\displaystyle\gamma_{2}^{\prime}(t_{0})\neq 0\,. (A.7)

Since γ1~((ai,bi))clBr(x)(x)=\tilde{\gamma_{1}}((a_{i},b_{i}))\cap\mathrm{cl}\,B_{r(x)}(x)=\varnothing and γ2([0,1])clBr(x)\gamma_{2}([0,1])\subset\mathrm{cl}\,B_{r}(x), we can choose γ2\gamma_{2} so that the concatenation of γ1\gamma_{1} and γ2\gamma_{2} defines a smooth embedding γ\gamma_{*} of SS1\SS^{1} into UiTU_{i}\subset T. Up to reparametrizing we may assume that γ(1)=x12\gamma_{*}(1)=x_{12}. Since γ1([0,1])V1\gamma_{1}([0,1])\subset V_{1} and V1(ST[s])=V_{1}\cap(S\cup T[s])=\varnothing, we have that

γ(SS1)(ST[s])=γ2([0,1])(ST[s])={x12}T[s]Br(x)(x).\gamma_{*}(\SS^{1})\cap(S\cup T[s])=\gamma_{2}([0,1])\cap(S\cup T[s])=\{x_{12}\}\subset T[s]\cap B_{r(x)}(x)\,. (A.8)

A first consequence of (A.8) is that γ(SS1)S=\gamma_{*}(\SS^{1})\cap S=\varnothing. Similarly, the curve γ:𝕊1Ω\gamma_{**}:\mathbb{S}^{1}\to\Omega defined via γ(t)=γ(t¯)\gamma_{**}(t)=\gamma_{*}(\overline{t}) (tSS1t\in\SS^{1}) where the bar denotes complex conjugation, has the same image as γ\gamma_{*} and thus satisfies γ(𝕊1)S=\gamma_{**}(\mathbb{S}^{1})\cap S=\varnothing as well. Therefore, in order to obtain a contradiction with |V2|=0|V_{2}|=0, it is enough to prove that either γ𝒞\gamma_{*}\in\mathcal{C} or γ𝒞\gamma_{**}\in\mathcal{C}. To this end we are now going to prove that one of γ\gamma_{*} or γ\gamma_{**} is homotopic to γ\gamma in TT (and thus in Ω\Omega), where γ\gamma is the curve from the tube (γ,Φ,T)𝒯(𝒞)(\gamma,\Phi,T)\in\mathcal{T}(\mathcal{C}) considered at the start of the argument.

Indeed, let 𝐩:SS1×B1nSS1\mathbf{p}:\SS^{1}\times B_{1}^{n}\to\SS^{1} denote the canonical projection 𝐩(t,x)=t\mathbf{p}(t,x)=t, and consider the curves σ=𝐩Φ1γ:SS1SS1\sigma_{*}=\mathbf{p}\circ\Phi^{-1}\circ\gamma_{*}:\SS^{1}\to\SS^{1} and σ=𝐩Φ1γ\sigma_{**}=\mathbf{p}\circ\Phi^{-1}\circ\gamma_{**}. By (A.8), σ1({s})={1}\sigma_{*}^{-1}(\{s\})=\{1\}, and 11 is a regular point of σ\sigma_{*} by (A.7) and since Φ\Phi is a diffeomorphism. Similarly, σ1({s})={1}\sigma_{**}^{-1}(\{s\})=\{1\} and 11 is a regular point of σ\sigma_{**}. Now by our construction of γ\gamma_{**}, exactly one of γ\gamma_{*} or γ\gamma_{**} is orientation preserving at 11 and the other is orientation reversing. So we may compute the winding numbers of σ\sigma_{*} and σ\sigma_{**} via (see e.g. [Mil97, pg 27]):

degσ=sgndetDσ(1)=sgndetDσ(1)=degσ{+1,1}.\displaystyle\mbox{deg}\,\sigma_{*}=\mbox{sgn}\,\det D\sigma_{*}(1)=-\mbox{sgn}\,\det D\sigma_{**}(1)=-\mbox{deg}\,\sigma_{**}\in\{+1,-1\}\,.

If we define σ=𝐩Φ1γ\sigma=\mathbf{p}\circ\Phi^{-1}\circ\gamma, then σ\sigma has winding number 11, and so is homotopic in 𝕊1\mathbb{S}^{1} to whichever of σ\sigma_{*} or σ\sigma_{**} has winding number 11. Since Φ\Phi is a diffeomorphism of SS1×B1n\SS^{1}\times B_{1}^{n} into Ω\Omega, we conclude that γ\gamma is homotopic relative to Ω\Omega to one of γ\gamma_{*} or γ\gamma_{**}, and, thus, that γ𝒞\gamma^{*}\in\mathcal{C} or γ𝒞\gamma_{**}\in\mathcal{C} as desired. ∎

Appendix B Convergence of every minimizing sequence of Ψbk(v)\Psi_{\rm bk}(v)

In proving Theorem 1.5 we have shown that every minimizing sequence {(Kj,Ej)}j\{(K_{j},E_{j})\}_{j} of Ψbk(v)\Psi_{\rm bk}(v) has a limit (K,E)(K,E) such that, denoting by B(w)B^{(w)} a ball of volume ww, it holds

Ψbk(v)=Ψbk(|E|)+P(B(v|E|)),Ψbk(|E|)=bk(K,E),\displaystyle\Psi_{\rm bk}(v)=\Psi_{\rm bk}(|E|)+P(B^{(v-|E|)})\,,\qquad\Psi_{\rm bk}(|E|)=\mathcal{F}_{\rm bk}(K,E)\,,

with both KK and EE bounded. In particular, minimizers of Ψbk(v)\Psi_{\rm bk}(v) can be constructed in the form (KB(v|E|)(x),EB(v|E|)(x))(K\cup\partial B^{(v-|E|)}(x),E\cup B^{(v-|E|)}(x)) provided xx is such that B(v|E|)(x)B^{(v-|E|)}(x) is disjoint from KE𝐖K\cup E\cup\mathbf{W}. This argument, although sufficient to prove the existence of minimizers of Ψbk(v)\Psi_{\rm bk}(v), it is not sufficient to prove the convergence of every minimizing sequence of Ψbk(v)\Psi_{\rm bk}(v), i.e., to exclude the possibility that |E|<v|E|<v. This is done in the following theorem at the cost of assuming the C2C^{2}-regularity of Ω\partial\Omega. This result will be important in the companion paper [MNR23a].

Theorem B.1.

If 𝐖\mathbf{W} is the closure of a bounded open set with C2C^{2}-boundary, 𝒞\mathcal{C} is a spanning class for 𝐖\mathbf{W}, and <\ell<\infty, then for every v>0v>0 and every minimizing sequence {(Kj,Ej)}j\{(K_{j},E_{j})\}_{j} of Ψbk(v)\Psi_{\rm bk}(v) there is a minimizer (K,E)(K,E) of Ψbk(v)\Psi_{\rm bk}(v) such that KK is n\mathcal{H}^{n}-rectifiable and, up to extracting subsequences and as jj\to\infty,

EjE,μjn  (ΩE)+2n  (KE(0)),E_{j}\to E\,,\qquad\mu_{j}\stackrel{{\scriptstyle\scriptscriptstyle{*}}}{{\rightharpoonup}}\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E)+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(K\cap E^{\scriptscriptstyle{(0)}})\,, (B.1)

where μj=n  (ΩEj)+2n  ((Kj)Ej(0))\mu_{j}=\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\Omega\cap\partial^{*}E_{j})+2\,\mathcal{H}^{n}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}(\mathcal{R}(K_{j})\cap E_{j}^{\scriptscriptstyle{(0)}}).

Proof.

By step three in the proof of Theorem 6.2, there is (K,E)𝒦B(K,E)\in\mathcal{K}_{\rm B} satisfying (B.1) and such that KK and EE are bounded, (K,E)(K,E) is a minimizer of Ψbk(|E|)\Psi_{\rm bk}(|E|), KK is n\mathcal{H}^{n}-rectifiable, and |E|v|E|\leq v; moreover, if v>|E|v>|E|, then there is xn+1x\in\mathbb{R}^{n+1} such that B(v|E|)(x)B^{(v-|E|)}(x) is disjoint from KE𝐖K\cup E\cup\mathbf{W} and (K,E)=(KB(v|E|)(x),EB(v|E|)(x))(K^{\prime},E^{\prime})=(K\cup\partial B^{(v-|E|)}(x),E\cup B^{(v-|E|)}(x)) is a minimizer of Ψbk(v)\Psi_{\rm bk}(v). We complete the proof by deriving a contradiction with the v=v|E|>0v^{*}=v-|E|>0 case. The idea is to relocate B(v)(x)B^{(v^{*})}(x) to save perimeter by touching 𝐖\partial\mathbf{W} or E\partial E; see Figure B.1.

First of all, we claim that K=ΩEK=\Omega\cap\partial E. If not, since (K,E)(K,E) and (K,E)(K^{\prime},E^{\prime}) respectively are minimizers of Ψbk(|E|)\Psi_{\rm bk}(|E|) and Ψbk(v)\Psi_{\rm bk}(v), then there are λ,λ\lambda,\lambda^{\prime}\in\mathbb{R} such that (K,E)(K,E) and (K,E)(K^{\prime},E^{\prime}) respectively satisfy (6.1) with λ\lambda and λ\lambda^{\prime}. By localizing (6.1) for (K,E)(K^{\prime},E^{\prime}) at points in ΩE\Omega\cap\partial^{*}E we see that it must be λ=λ\lambda=\lambda^{\prime}; by localizing at points in B(v|E|)(x)\partial B^{(v-|E|)}(x), we see that λ\lambda is equal to the mean curvature of B(v|E|)(x)\partial B^{(v-|E|)}(x), so that λ>0\lambda>0; by arguing as in the proof of [KMS21, Theorem 2.9] (see [Nov23] for the details), we see that if K(ΩE)K\setminus(\Omega\cap\partial E)\neq\varnothing, then λ0\lambda\leq 0, a contradiction.

Having established that K=ΩEK=\Omega\cap\partial E, we move an half-space HH compactly containing cl(E)𝐖\mathrm{cl}\,(E)\cup\mathbf{W} until the boundary hyperplane H\partial H first touches cl(E)𝐖\mathrm{cl}\,(E)\cup\mathbf{W}. Up to rotation and translation, we can thus assume that H={xn+1>0}H=\{x_{n+1}>0\} and

0cl(E)𝐖cl(H).0\in\mathrm{cl}\,(E)\cup\mathbf{W}\subset\mathrm{cl}\,(H)\,. (B.2)

We split (B.2) into two cases, 0ΩE0\in\Omega\cap\partial E and 0𝐖0\in\mathbf{W}, that are then separately discussed for the sake of clarity. In both cases we set x=(x,xn+1)n×n+1x=(x^{\prime},x_{n+1})\in\mathbb{R}^{n}\times\mathbb{R}\equiv\mathbb{R}^{n+1}, and set

𝐂δ\displaystyle{\bf C}_{\delta} =\displaystyle= {x:xn+1(0,δ),|x|<δ},\displaystyle\{x:x_{n+1}\in(0,\delta)\,,|x^{\prime}|<\delta\}\,,
𝐋δ\displaystyle{\bf L}_{\delta} =\displaystyle= {x:|x|=δ,xn+1(0,δ)},\displaystyle\{x:|x^{\prime}|=\delta,x_{n+1}\in(0,\delta)\}\,,
𝐓δ\displaystyle{\bf T}_{\delta} =\displaystyle= {x:xn+1=δ,|x|<δ},\displaystyle\{x:x_{n+1}=\delta\,,|x^{\prime}|<\delta\}\,,
𝐃δ\displaystyle{\bf D}_{\delta} =\displaystyle= {x:xn+1=0,|x|<δ},\displaystyle\{x:x_{n+1}=0\,,|x^{\prime}|<\delta\}\,,

for every δ>0\delta>0.

Case one, 0ΩE0\in\Omega\cap\partial E: In this case, by the maximum principle [DM19, Lemma 3], (6.1), and the Allard regularity theorem, we can find δ0>0\delta_{0}>0 and uC2(𝐃δ0;[0,δ0])u\in C^{2}({\bf D}_{\delta_{0}};[0,\delta_{0}]) with u(0)=0u(0)=0 and u(0)=0\nabla u(0)=0 such that 𝐂δ0Ω{\bf C}_{\delta_{0}}\subset\!\subset\Omega and

E𝐂δ0={x𝐂δ0:δ0>xn+1>u(x)},\displaystyle E\cap{\bf C}_{\delta_{0}}=\big{\{}x\in{\bf C}_{\delta_{0}}:\delta_{0}>x_{n+1}>u(x^{\prime})\big{\}}\,, (B.3)
(E)𝐂δ0={x𝐂δ0:xn+1=u(x)}.\displaystyle(\partial E)\cap{\bf C}_{\delta_{0}}=\big{\{}x\in{\bf C}_{\delta_{0}}:x_{n+1}=u(x^{\prime})\big{\}}\,.

Since 0u(x)C|x|20\leq u(x^{\prime})\leq C\,|x^{\prime}|^{2} for some C=C(E)C=C(E), if we set

Γδ={x𝐂δ:0<xn+1<u(x)},δ(0,δ0),\Gamma_{\delta}=\Big{\{}x\in{\bf C}_{\delta}:0<x_{n+1}<u(x^{\prime})\Big{\}}\,,\qquad\delta\in(0,\delta_{0})\,, (B.4)

then we have

|Γδ|\displaystyle|\Gamma_{\delta}|\!\! \displaystyle\leq Cδn+2,\displaystyle\!\!C\,\delta^{n+2}\,, (B.5)
P(Γδ;𝐋δ)\displaystyle P\big{(}\Gamma_{\delta};{\bf L}_{\delta}\big{)}\!\! \displaystyle\leq Cδn+1.\displaystyle\!\!C\,\delta^{n+1}\,. (B.6)

We then set

Eδ=EΓδ(Brδ(zδ)H),E_{\delta}=E\cup\Gamma_{\delta}\cup\big{(}B_{r_{\delta}}(z_{\delta})\setminus H\big{)}\,, (B.7)

see

Refer to caption
HHEδE_{\delta}δ\deltaEδE_{\delta}Γδ\Gamma_{\delta}EEEEδ\delta0rδr_{\delta}zδz_{\delta}(a)(a)𝐖\mathbf{W}(b)(b)
Figure B.1. (a): the construction of EδE_{\delta} when 0ΩE0\in\Omega\cap\partial E; (b) the construction of EδE_{\delta} when 0𝐖0\in\mathbf{W}.

Figure B.1-(a), where rδ>0r_{\delta}>0 and zδn+1cl(H)z_{\delta}\in\mathbb{R}^{n+1}\setminus\mathrm{cl}\,(H) are uniquely determined by requiring that, first,

cl(Brδ(zδ))H=𝐂δH={x:xn+1=0,|x|δ},\displaystyle\mathrm{cl}\,(B_{r_{\delta}}(z_{\delta}))\cap\partial H=\partial{\bf C}_{\delta}\cap\partial H=\big{\{}x:x_{n+1}=0\,,|x^{\prime}|\leq\delta\big{\}}\,, (B.8)

and, second, that

|Eδ|=v.|E_{\delta}|=v\,. (B.9)

To see that this choice is possible, we first notice that, since EΓδ=E\cap\Gamma_{\delta}=\varnothing, (B.9) is equivalent to

|Brδ(zδ)H|=v|E||Γδ|=v|Γδ|.\big{|}B_{r_{\delta}}(z_{\delta})\setminus H\big{|}=v-|E|-|\Gamma_{\delta}|=v^{*}-|\Gamma_{\delta}|\,. (B.10)

Taking (B.5) into account we see that (B.8) and (B.10) uniquely determine zδn+1z_{\delta}\in\mathbb{R}^{n+1} and rδ>0r_{\delta}>0 as soon as δ0\delta_{0} is small enough to guarantee v|Γδ0|>0v^{*}-|\Gamma_{\delta_{0}}|>0. In fact, by (B.5), v|Γδ|v>0v^{*}-|\Gamma_{\delta}|\to v^{*}>0 with n(𝐂δH)0\mathcal{H}^{n}(\partial{\bf C}_{\delta}\cap\partial H)\to 0 as δ0+\delta\to 0^{+}, so that, up to further decrease δ0\delta_{0}, we definitely have zδHz_{\delta}\not\in H, and

|rδ(vωn+1)1/(n+1)|Cδn+2,\Big{|}r_{\delta}-\Big{(}\frac{v^{*}}{\omega_{n+1}}\Big{)}^{1/(n+1)}\Big{|}\leq C\,\delta^{n+2}\,, (B.11)

where C=C(E,n,v)C=C(E,n,v^{*}).

We now use the facts that KE(1)K\cup E^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} and that EEδE\subset E_{\delta} to prove that

(Kδ,Eδ)=((ΩEδ)(KEδ(0)),Eδ)(K_{\delta},E_{\delta})=((\Omega\cap\partial^{*}E_{\delta})\cup(K\cap E_{\delta}^{\scriptscriptstyle{(0)}}),E_{\delta}) (B.12)

is such that KδEδ(1)K_{\delta}\cup E_{\delta}^{\scriptscriptstyle{(1)}} is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W} (and thus is admissible in Ψbk(v)\Psi_{\rm bk}(v) by (B.9)). To this end, it is enough to show that

KE(1)nKδEδ(1).\displaystyle K\cup E^{\scriptscriptstyle{(1)}}\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}K_{\delta}\cup E_{\delta}^{\scriptscriptstyle{(1)}}\,. (B.13)

Indeed, by EEδE\subset E_{\delta} and Federer’s theorem (1.37) we have

E(1)Eδ(1),Eδ(0)E(0),E(1)EnEδ(1)Eδ.\displaystyle E^{\scriptscriptstyle{(1)}}\subset E_{\delta}^{\scriptscriptstyle{(1)}}\,,\qquad E_{\delta}^{\scriptscriptstyle{(0)}}\subset E^{\scriptscriptstyle{(0)}}\,,\qquad E^{\scriptscriptstyle{(1)}}\cup\partial^{*}E\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}E_{\delta}^{\scriptscriptstyle{(1)}}\cup\partial^{*}E_{\delta}\,. (B.14)

(Notice indeed that EE(1/2)n+1Eδ(0)\partial^{*}E\subset E^{\scriptscriptstyle{(1/2)}}\subset\mathbb{R}^{n+1}\setminus E_{\delta}^{\scriptscriptstyle{(0)}}). Next, using in order Federer’s theorem (1.37), (B.14) and KΩK\subset\Omega, and the definition of KδK_{\delta}, we have

E(1)(KEδ(0))=nE(1)[K(EδEδ(1))]nEδ(1)(ΩEδ)Eδ(1)Kδ.E^{\scriptscriptstyle{(1)}}\cup(K\setminus E_{\delta}^{\scriptscriptstyle{(0)}})\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{=}E^{\scriptscriptstyle{(1)}}\cup[K\cap(\partial^{*}E_{\delta}\cup E_{\delta}^{\scriptscriptstyle{(1)}})]\overset{\scriptscriptstyle{\mathcal{H}^{n}}}{\subset}E_{\delta}^{\scriptscriptstyle{(1)}}\cup(\Omega\cap\partial^{*}E_{\delta})\subset E_{\delta}^{\scriptscriptstyle{(1)}}\cup K_{\delta}\,.

But KEδ(0)KδK\cap E_{\delta}^{\scriptscriptstyle{(0)}}\subset K_{\delta} by definition, which combined with the preceding containment completes the proof of (B.13). Having proved that (Kδ,Eδ)(K_{\delta},E_{\delta}) is admissible in Ψbk(v)\Psi_{\rm bk}(v), we have

bk(K,E)+P(B(v))=Ψbk(v)bk(Kδ,Eδ).\mathcal{F}_{\rm bk}(K,E)+P(B^{(v^{*})})=\Psi_{\rm bk}(v)\leq\mathcal{F}_{\rm bk}(K_{\delta},E_{\delta})\,. (B.15)

By (B.15), the definition of KδK_{\delta}, and (B.14), we find

P(E;Ω)+2n(KE(0))+P(B(v))P(Eδ;Ω)+2n(KδEδ(0))\displaystyle P(E;\Omega)+2\,\mathcal{H}^{n}(K\cap E^{\scriptscriptstyle{(0)}})+P(B^{(v^{*})})\leq P(E_{\delta};\Omega)+2\,\mathcal{H}^{n}(K_{\delta}\cap E_{\delta}^{\scriptscriptstyle{(0)}})
P(Eδ;Ω)+2n(KEδ(0))P(Eδ;Ω)+2n(KE(0)),\displaystyle\hskip 39.83368pt\leq P(E_{\delta};\Omega)+2\,\mathcal{H}^{n}(K\cap E_{\delta}^{\scriptscriptstyle{(0)}})\leq P(E_{\delta};\Omega)+2\,\mathcal{H}^{n}(K\cap E^{\scriptscriptstyle{(0)}})\,,

from which we deduce

P(E;Ω)+P(B(v))P(Eδ;Ω).\displaystyle P(E;\Omega)+P(B^{(v^{*})})\leq P(E_{\delta};\Omega)\,. (B.16)

We now notice that EδE_{\delta} coincides with EE in the open set ΩHcl(𝐂δ)\Omega\cap H\setminus\mathrm{cl}\,({\bf C}_{\delta}), and with Brδ(zδ)B_{r_{\delta}}(z_{\delta}) in the open set n+1cl(H)\mathbb{R}^{n+1}\setminus\mathrm{cl}\,(H), so that

(ΩHcl(𝐂δ))Eδ=(ΩHcl(𝐂δ))E,\displaystyle\Big{(}\Omega\cap H\setminus\mathrm{cl}\,({\bf C}_{\delta})\Big{)}\cap\partial^{*}E_{\delta}=\Big{(}\Omega\cap H\setminus\mathrm{cl}\,({\bf C}_{\delta})\Big{)}\cap\partial^{*}E\,,
(Ωcl(H))Eδ=(Brδ(zδ))cl(H),\displaystyle\big{(}\Omega\setminus\mathrm{cl}\,(H)\big{)}\cap\partial^{*}E_{\delta}=\big{(}\partial B_{r_{\delta}}(z_{\delta})\big{)}\setminus\mathrm{cl}\,(H)\,,

and (B.16) is equivalent to

P(E;Ω(Hcl(𝐂δ))+P(B(v))\displaystyle P\big{(}E;\Omega\cap(\partial H\cup\mathrm{cl}\,({\bf C}_{\delta})\big{)}+P(B^{(v^{*})}) (B.17)
P(Eδ;Ω(Hcl(𝐂δ))+P(Brδ(zδ);n+1cl(H)).\displaystyle\leq P\big{(}E_{\delta};\Omega\cap(\partial H\cup\mathrm{cl}\,({\bf C}_{\delta})\big{)}+P(B_{r_{\delta}}(z_{\delta});\mathbb{R}^{n+1}\setminus\mathrm{cl}\,(H))\,.

In fact, it is easily proved that (E)(H)cl(𝐂δ)=(Eδ)(H)cl(𝐂δ)(\partial^{*}E)\cap(\partial H)\setminus\mathrm{cl}\,({\bf C}_{\delta})=(\partial^{*}E_{\delta})\cap(\partial H)\setminus\mathrm{cl}\,({\bf C}_{\delta}) (which is evident from Figure B.1), so that (B.17) readily implies

P(B(v))P(Eδ;Ωcl(𝐂δ))+P(Brδ(zδ);n+1cl(H)).\displaystyle P(B^{(v^{*})})\leq P\big{(}E_{\delta};\Omega\cap\mathrm{cl}\,({\bf C}_{\delta})\big{)}+P(B_{r_{\delta}}(z_{\delta});\mathbb{R}^{n+1}\setminus\mathrm{cl}\,(H))\,. (B.18)

Now, 𝐂δΩ{\bf C}_{\delta}\subset\!\subset\Omega. Moreover, by (B.3), we have that 𝐓δ{\bf T}_{\delta} (the top part of 𝐂δ\partial{\bf C}_{\delta}) is contained in E(1)Eδ(1)E^{\scriptscriptstyle{(1)}}\subset E_{\delta}^{\scriptscriptstyle{(1)}}, and is thus n\mathcal{H}^{n}-disjoint from Eδ\partial^{*}E_{\delta}. Similarly, again by (B.3) we have EΓδ=𝐂δE\cup\Gamma_{\delta}={\bf C}_{\delta}, and thus 𝐃δ(EΓδ)(1/2){\bf D}_{\delta}\subset(E\cup\Gamma_{\delta})^{\scriptscriptstyle{(1/2)}}; at the same time, by (B.8) we have 𝐃δ(Brδ(zδ)H)(1/2){\bf D}_{\delta}\subset(B_{r_{\delta}}(z_{\delta})\setminus H)^{\scriptscriptstyle{(1/2)}}; therefore 𝐃δEδ(1){\bf D}_{\delta}\subset E_{\delta}^{\scriptscriptstyle{(1)}}, and thus 𝐃δ{\bf D}_{\delta} is n\mathcal{H}^{n}-disjoint from Eδ\partial^{*}E_{\delta}. Finally, again by EΓδ=𝐂δE\cup\Gamma_{\delta}={\bf C}_{\delta} we see that P(Eδ;𝐂δ)=0P(E_{\delta};{\bf C}_{\delta})=0. Therefore, in conclusion,

P(Eδ;Ωcl(𝐂δ))=P(Eδ;𝐋δ)=P(Γδ;𝐋δ)Cδn+1,\displaystyle P\big{(}E_{\delta};\Omega\cap\mathrm{cl}\,({\bf C}_{\delta})\big{)}=P(E_{\delta};{\bf L}_{\delta})=P(\Gamma_{\delta};{\bf L}_{\delta})\leq C\,\delta^{n+1}\,, (B.19)

where we have used again first (B.3), and then (B.6). Combining (B.18)-(B.19) we get

P(B(v))P(Brδ(zδ);n+1cl(H))+Cδn+1.\displaystyle P(B^{(v^{*})})\leq P(B_{r_{\delta}}(z_{\delta});\mathbb{R}^{n+1}\setminus\mathrm{cl}\,(H))+C\,\delta^{n+1}\,. (B.20)

Finally, by (B.8), (B.5), and (B.11) we have

P(Brδ(zδ);n+1cl(H))P(B(v))C(n)δn;P(B_{r_{\delta}}(z_{\delta});\mathbb{R}^{n+1}\setminus\mathrm{cl}\,(H))\leq P(B^{(v^{*})})-C(n)\,\delta^{n}\,;

by combining this estimate with (B.20), we reach a contradiction for δ\delta small enough.

Case two, 0𝐖0\in\mathbf{W}: In this case, by the C2C^{2}-regularity of Ω\partial\Omega we can find δ0>0\delta_{0}>0 and uC2(𝐃δ0;[0,δ0])u\in C^{2}({\bf D}_{\delta_{0}};[0,\delta_{0}]) with u(0)=0u(0)=0 and u(0)=0\nabla u(0)=0 such that

𝐖𝐂δ0={x𝐂δ0:δ0>xn+1>u(x)},\displaystyle\mathbf{W}\cap{\bf C}_{\delta_{0}}=\big{\{}x\in{\bf C}_{\delta_{0}}:\delta_{0}>x_{n+1}>u(x^{\prime})\big{\}}\,, (B.21)
(Ω)𝐂δ={x𝐂δ0:xn+1=u(x)}.\displaystyle(\partial\Omega)\cap{\bf C}_{\delta}=\big{\{}x\in{\bf C}_{\delta_{0}}:x_{n+1}=u(x^{\prime})\big{\}}\,.

We have 0u(x)C|x|20\leq u(x^{\prime})\leq C\,|x^{\prime}|^{2} for every |x|<δ0|x^{\prime}|<\delta_{0} (and some C=C(𝐖)C=C(\mathbf{W})), so that defining Γδ\Gamma_{\delta} as in (B.4) we still obtain (B.5) and (B.6). We then define EδE_{\delta}, rδr_{\delta}, and zδz_{\delta}, as in (B.7), (B.8) and (B.9). Notice that now EE and Γδ\Gamma_{\delta} may not be disjoint (see Figure B.1-(b)), therefore (B.9) is not equivalent to (B.10), but to

|Brδ(zδ)H|=v|E||ΓδE|=v|ΓδE|.\big{|}B_{r_{\delta}}(z_{\delta})\setminus H\big{|}=v-|E|-|\Gamma_{\delta}\setminus E|=v^{*}-|\Gamma_{\delta}\setminus E|\,.

This is still sufficient to repeat the considerations based on (B.8) and (B.5) proving that rδr_{\delta} and zδz_{\delta} are uniquely determined, and satisfy (B.11). We can repeat the proof that (Kδ,Eδ)(K_{\delta},E_{\delta}) defined as in (B.12) is admissible in Ψbk(v)\Psi_{\rm bk}(v) (since that proof was based only on the inclusion EEδE\subset E_{\delta}), and thus obtain (B.16). The same considerations leading from (B.16) to (B.18) apply in the present case too, and so we land on

P(B(v))P(Eδ;Ωcl(𝐂δ))+P(Brδ(zδ);n+1cl(H)).\displaystyle P(B^{(v^{*})})\leq P\big{(}E_{\delta};\Omega\cap\mathrm{cl}\,({\bf C}_{\delta})\big{)}+P(B_{r_{\delta}}(z_{\delta});\mathbb{R}^{n+1}\setminus\mathrm{cl}\,(H))\,. (B.22)

Now, by (B.21), 𝐓δ{\bf T}_{\delta} is contained in 𝐖\mathbf{W}, so that P(Eδ;𝐓δ)=0P(E_{\delta};{\bf T}_{\delta})=0. At the same time, if x=(x,0)𝐃δΩx=(x^{\prime},0)\in{\bf D}_{\delta}\cap\Omega, then u(x)>0u(x^{\prime})>0, and thus x(EδH)(1/2)x\in(E_{\delta}\cap H)^{\scriptscriptstyle{(1/2)}}; since, by (B.8), we also have x(EδH)(1/2)x\in(E_{\delta}\setminus H)^{\scriptscriptstyle{(1/2)}}, we conclude that 𝐃δΩEδ(1){\bf D}_{\delta}\cap\Omega\subset E_{\delta}^{\scriptscriptstyle{(1)}}, and thus that

P(Eδ;Ωcl(𝐂δ))=P(Eδ;Ω𝐋δ)n(Ω𝐋δ)Cδn+1,P\big{(}E_{\delta};\Omega\cap\mathrm{cl}\,({\bf C}_{\delta})\big{)}=P\big{(}E_{\delta};\Omega\cap{\bf L}_{\delta}\big{)}\leq\mathcal{H}^{n}(\Omega\cap{\bf L}_{\delta})\leq C\,\delta^{n+1}\,,

where we have used 0u(x)C|x|20\leq u(x^{\prime})\leq C\,|x^{\prime}|^{2} for every |x|<δ0|x^{\prime}|<\delta_{0} again. We thus deduce from (B.22) that

P(B(v))P(Brδ(zδ);n+1cl(H))+Cδn+1,P(B^{(v^{*})})\leq P(B_{r_{\delta}}(z_{\delta});\mathbb{R}^{n+1}\setminus\mathrm{cl}\,(H))+C\,\delta^{n+1}\,,

and from here we conclude as in case one. ∎

Appendix C An elementary lemma

In this appendix we provide a proof of Lemma 7.2. The proof is an immediate corollary of a geometric property of closed 𝒞\mathcal{C}-spanning sets (see (C.2)-(C.3) below) first proved in n+1\mathbb{R}^{n+1} for n2n\geq 2 [DLDRG19, Lemma 4.1]. Here we extend this property to the plane. The difference between 2\mathbb{R}^{2} and n+1\mathbb{R}^{n+1} for n2n\geq 2 stems from a part of the argument where one constructs a new admissible spanning curve by modifying an existing one inside a ball. Specifically, ensuring that the new curve does not intersect itself requires an extra argument in 2\mathbb{R}^{2}.

Lemma C.1.

Let n1n\geq 1, 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be closed, 𝒞\mathcal{C} be a spanning class for 𝐖\mathbf{W}, SΩ:=n+1𝐖S\subset\Omega:=\mathbb{R}^{n+1}\setminus\mathbf{W} be relatively closed and 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, and Br(x)ΩB_{r}(x)\subset\!\subset\Omega. Let {Γi}i\{\Gamma_{i}\}_{i} be the countable family of equivalence classes of Br(x)S\partial B_{r}(x)\setminus S determined by the relation:

yxγ~C0([0,1],clBr(x)S):γ~(0)=yγ~(1)=zγ~((0,1))Br(x).\displaystyle y\sim x\iff\mbox{$\exists\tilde{\gamma}\in C^{0}([0,1],\mathrm{cl}\,B_{r}(x)\setminus S):\tilde{\gamma}(0)=y$, $\tilde{\gamma}(1)=z$, $\tilde{\gamma}((0,1))\subset B_{r}(x)$}\,. (C.1)

Then if γ𝒞\gamma\in\mathcal{C}, either

γ(SBr(x))\displaystyle\gamma\cap(S\setminus B_{r}(x))\neq\emptyset (C.2)

or there exists a connected component σ\sigma of γclBr(x)\gamma\cap\mathrm{cl}\,B_{r}(x) which is homeomorphic to an interval and such that

the endpoints of σ\sigma belong to two distinct equivalence classes of Br(x)S\partial B_{r}(x)\setminus S. (C.3)

In particular, the conclusion of Lemma 7.2 holds.

Remark C.2.

The planar version of Lemma C.1 allows one to extend the main existence result [DLDRG19, Theorem 2.7] to 2\mathbb{R}^{2}.

Proof of Lemma C.1.

The proof is divided into two pieces. First we show how to deduce Lemma 7.2 from the fact that at least one of (C.2)-(C.3) holds. Then we show in 2\mathbb{R}^{2} that (C.3) must hold whenever (C.2) does not, completing the lemma since the case n2n\geq 2 is contained in [DLDRG19, Lemma 4.1].

Conclusion of Lemma 7.2 from (C.2)-(C.3): We must show that either γ(𝕊1)Br(x)\gamma(\mathbb{S}^{1})\setminus B_{r}(x)\neq\varnothing or that it intersects at least two open connected components of Br(x)SB_{r}(x)\setminus S. If γ(𝕊1)Br(x)\gamma(\mathbb{S}^{1})\setminus B_{r}(x)\neq\varnothing we are done, so suppose that γ(𝕊1)Br(x)=\gamma(\mathbb{S}^{1})\setminus B_{r}(x)=\varnothing. Then (C.3) must be true, so that the endpoints of some arc σ=γ((a,b))Br(x)\sigma=\gamma((a,b))\subset B_{r}(x) for an interval (a,b)SS1(a,b)\subset\SS^{1} belong to distinct equivalence classes. Choose ρ\rho small enough so that Bρ(γ(a))Bρ(γ(b))ΩSB_{\rho}(\gamma(a))\cup B_{\rho}(\gamma(b))\subset\Omega\setminus S and aa^{\prime}, bIb^{\prime}\in I such that γ(a)Bρ(γ(a))\gamma(a^{\prime})\in B_{\rho}(\gamma(a)) and γ(b)Bρ(γ(b))\gamma(b^{\prime})\in B_{\rho}(\gamma(b)). If γ(a)\gamma(a^{\prime}) and γ(b)\gamma(b^{\prime}) belonged to the same open connected component of Br(x)SB_{r}(x)\setminus S, we would contradict (C.3), so they belong to different components as desired.

Verification of (C.2)-(C.3) in 2\mathbb{R}^{2}: As in [DLGM17a, Lemma 10], we may reduce to the case where γ\gamma intersects Br(x)\partial B_{r}(x) transversally at finitely many points {γ(ak)}k=1K{γ(bk)}k=1K\{\gamma(a_{k})\}_{k=1}^{K}\cup\{\gamma(b_{k})\}_{k=1}^{K} such that γBr(x)=kγ((ak,bk))\gamma\cap B_{r}(x)=\cup_{k}\gamma((a_{k},b_{k})) and {[ak,bk]}k\{[a_{k},b_{k}]\}_{k} are mutually disjoint closed arcs in 𝕊1\mathbb{S}^{1}. If (C.2) holds we are done, so we assume that

γSBr(x)=\displaystyle\gamma\cap S\setminus B_{r}(x)=\varnothing (C.4)

and prove (C.3). Note that each pair {γ(ak),γ(bk)}\{\gamma(a_{k}),\gamma(b_{k})\} bounds two open arcs in Br(x)\partial B_{r}(x); we make a choice now as follows. Choose s0Br(x)k{γ(ak),γ(bk)}s_{0}\in\partial B_{r}(x)\setminus\cup_{k}\{\gamma(a_{k}),\gamma(b_{k})\}. Based on our choice of s0s_{0}, for each kk there is a unique open arc kBr(x)\ell_{k}\subset\partial B_{r}(x) such that Br(x)k={γ(ak),γ(bk)}\partial_{\partial B_{r}(x)}\ell_{k}=\{\gamma(a_{k}),\gamma(b_{k})\} and s0clBr(x)ks_{0}\notin\mathrm{cl}\,_{\partial B_{r}(x)}\ell_{k}. We claim that

if kk, then either kk or kk.\displaystyle\mbox{if $k\neq k^{\prime}$, then either $\ell_{k}\subset\!\subset\ell_{k^{\prime}}$ or $\ell_{k^{\prime}}\subset\!\subset\ell_{k}$}\,. (C.5)

To prove (C.5): We consider simple closed curves γk\gamma_{k} with images γ((ak,bk))clBr(x)k\gamma((a_{k},b_{k}))\cup\mathrm{cl}\,_{\partial B_{r}(x)}\ell_{k}. By the Jordan curve theorem, each γk\gamma_{k} defines a connected open subset UkU_{k} of Br(x)B_{r}(x) with UkBr(x)=clBr(x)k\partial U_{k}\cap\partial B_{r}(x)=\mathrm{cl}\,_{\partial B_{r}(x)}\ell_{k}. Aiming for a contradiction, if (C.5) were false, then for some kkk\neq k^{\prime}, either

γ(ak)kclUk\gamma(a_{k})\in\ell_{k^{\prime}}\subset\mathrm{cl}\,U_{k^{\prime}} and γ(bk)Br(x)clBr(x)kBr(x)clUk\gamma(b_{k})\in\partial B_{r}(x)\setminus\mathrm{cl}\,_{\partial B_{r}(x)}\ell_{k^{\prime}}\subset\partial B_{r}(x)\setminus\mathrm{cl}\,U_{k^{\prime}} or
γ(bk)kclUk and γ(ak)Br(x)clBr(x)kBr(x)clUk;\displaystyle\mbox{$\gamma(b_{k})\in\ell_{k^{\prime}}\subset\mathrm{cl}\,U_{k^{\prime}}$ and $\gamma(a_{k})\in\partial B_{r}(x)\setminus\mathrm{cl}\,_{\partial B_{r}(x)}\ell_{k^{\prime}}\subset\partial B_{r}(x)\setminus\mathrm{cl}\,U_{k^{\prime}}$}\,;

in particular, γ((ak,bk))\gamma((a_{k},b_{k})) has non-trivial intersection with both the open sets UkU_{k^{\prime}} and Br(x)clUkB_{r}(x)\setminus\mathrm{cl}\,U_{k^{\prime}}. By the continuity of γ\gamma and the connectedness of (ak,bk)(a_{k},b_{k}), we thus deduce that γ((ak,bk))Uk\gamma((a_{k},b_{k}))\cap\partial U_{k^{\prime}}\neq\varnothing. Upon recalling that γ((ak,bk))Br(x)\gamma((a_{k},b_{k}))\subset B_{r}(x), we find γ((ak,bk))UkBr(x)=γ((ak,bk))γ((ak,bk))\gamma((a_{k},b_{k}))\cap\partial U_{k^{\prime}}\cap B_{r}(x)=\gamma((a_{k},b_{k}))\cap\gamma((a_{k^{\prime}},b_{k^{\prime}}))\neq\varnothing. But this contradicts the fact that γ\gamma smoothly embeds SS1\SS^{1} into Ω\Omega. The proof of (C.5) is finished.

Returning to the proof of (C.3), let us assume for contradiction that

γ(ak)γ(bk)1kK.\displaystyle\gamma(a_{k})\sim\gamma(b_{k})\quad\forall 1\leq k\leq K\,. (C.6)

We are going to use (C.4), (C.5), and (C.6) to create a piecewise smooth embedding γ¯:𝕊1Ω\overline{\gamma}:\mathbb{S}^{1}\to\Omega which is a homotopic deformation of γ\gamma (and thus approximable by elements in 𝒞\mathcal{C}) such that γ¯S=\overline{\gamma}\cap S=\varnothing. After reindexing the equivalence classes Γi\Gamma_{i}, we may assume that {Γ1,,ΓIγ}\{\Gamma_{1},\dots,\Gamma_{I_{\gamma}}\} are those equivalence classes containing any pair {γ(ak),γ(bk)}\{\gamma(a_{k}),\gamma(b_{k})\} for 1kK1\leq k\leq K. We will construct γ¯\overline{\gamma} in steps by redefining γ\gamma on those [ak,bk][a_{k},b_{k}] with images under γ\gamma having endpoints belonging to the same Γi\Gamma_{i}. For future use, let Ωi\Omega_{i} be the equivalence classes of Br(x)SB_{r}(x)\setminus S determined by the relation (C.1). Note that they are open connected components of Br(x)SB_{r}(x)\setminus S.

Construction corresponding to Γ1\Gamma_{1}: Relabelling in kk if necessary, we may assume that {1,,K1}\{1,\dots,K_{1}\} for some 1K1K1\leq K_{1}\leq K are the indices such that {γ(ak),γ(bk)}Γ1\{\gamma(a_{k}),\gamma(b_{k})\}\subset\Gamma_{1}. By further relabelling and applying (C.5) we may assume: first, that 1\ell_{1} is a “maximal” arc among {1,,K1}\{\ell_{1},\dots,\ell_{K_{1}}\}, in other words

for given k{2,K1}, either 1k= or k1;\displaystyle\mbox{for given $k\in\{2,\dots K_{1}\}$, either $\ell_{1}\cap\ell_{k}=\varnothing$ or $\ell_{k}\subset\!\subset\ell_{1}$}\,; (C.7)

and second, that for some K11K1K_{1}^{1}\leq K_{1}, {2,,K11}\{\ell_{2},\dots,\ell_{K_{1}^{1}}\} are those arcs contained in 1\ell_{1}. Since Ω1\Omega_{1} is open and connected, we may connect γ(a1)\gamma(a_{1}) to γ(b1)\gamma(b_{1}) by a smooth embedding γ¯1:[a1,b1]clBr(x)S\overline{\gamma}_{1}:[a_{1},b_{1}]\to\mathrm{cl}\,B_{r}(x)\setminus S with γ¯1((a1,b1))Ω1\overline{\gamma}_{1}((a_{1},b_{1}))\subset\Omega_{1}. Also, by the Jordan curve theorem, 1γ¯1\ell_{1}\cup\overline{\gamma}_{1} defines an open connected subset W1W_{1} of Br(x)B_{r}(x) with W1S=\partial W_{1}\cap S=\varnothing. Using (C.5), we now argue towards constructing pairwise disjoint smooth embeddings γ¯k:[ak,bk]Γ1Ω1\overline{\gamma}_{k}:[a_{k},b_{k}]\to\Gamma_{1}\cup\Omega_{1}.

We first claim that

W1S is path-connected.\displaystyle\mbox{$W_{1}\setminus S$ is path-connected}\,. (C.8)

To prove (C.8), consider any y,zW1Sy,z\in W_{1}\setminus S. Since Ω1W1S\Omega_{1}\supset W_{1}\setminus S is open and path-connected, we may obtain continuous γ~:[0,1]Ω1\tilde{\gamma}:[0,1]\to\Omega_{1} connecting yy and zz. If γ~([0,1])W1S\tilde{\gamma}([0,1])\subset W_{1}\setminus S, we are done. Otherwise, γ~(Ω1(W1S))=Ω1W1\varnothing\neq\tilde{\gamma}\cap(\Omega_{1}\setminus(W_{1}\setminus S))=\Omega_{1}\setminus W_{1}, with the equality following from Ω1S=\Omega_{1}\cap S=\varnothing. Combining this information with γ~({0,1})W1S\tilde{\gamma}(\{0,1\})\subset W_{1}\setminus S, we may therefore choose [δ1,δ2](0,1)[\delta_{1},\delta_{2}]\subset(0,1) to be the smallest interval such that γ~([0,1][δ1,δ2])W1S\tilde{\gamma}([0,1]\setminus[\delta_{1},\delta_{2}])\subset W_{1}\setminus S. On (δ1,δ2)(\delta_{1},\delta_{2}), we redefine γ~\tilde{\gamma} using the fact that γ~({δ1,δ2})W1Br(x)=γ¯1((a1,b1))\tilde{\gamma}(\{\delta_{1},\delta_{2}\})\subset\partial W_{1}\cap B_{r}(x)=\overline{\gamma}_{1}((a_{1},b_{1})) by letting γ~((δ1,δ2))=γ¯1(I)\tilde{\gamma}((\delta_{1},\delta_{2}))=\overline{\gamma}_{1}(I), where γ¯1(I)\overline{\gamma}_{1}(I) has endpoints γ~(δ1)\tilde{\gamma}(\delta_{1}) and γ~(δ2)\tilde{\gamma}(\delta_{2}) and I(a1,b1)I\subset(a_{1},b_{1}). The modified γ~\tilde{\gamma} is a concatenation of continuous curves and is thus continuous; furthermore, γ~1(W1S)=[0,δ1)(δ2,1]\tilde{\gamma}^{-1}(W_{1}\setminus S)=[0,\delta_{1})\cup(\delta_{2},1]. It only remains to “push” γ~\tilde{\gamma} entirely inside W1SW_{1}\setminus S, which we may easily achieve by projecting γ~((δ1ε,δ2+ε))\tilde{\gamma}((\delta_{1}-\varepsilon,\delta_{2}+\varepsilon)) inside W1SW_{1}\setminus S for small ε\varepsilon using the distance function to the smooth curve γ¯1(a1,b1)=W1Br(x)Br(x)S\overline{\gamma}_{1}(a_{1},b_{1})=\partial W_{1}\cap B_{r}(x)\subset B_{r}(x)\setminus S. This completes (C.8).

But now since W1SW_{1}\setminus S is path-connected and open, we may connect any two points in it by a smooth embedding of [0,1][0,1], which in particular allows us to connect γ(a2)\gamma(a_{2}) and γ(b2)\gamma(b_{2}) by smooth embedding γ¯2:[a2,b2]clW1S\overline{\gamma}_{2}:[a_{2},b_{2}]\to\mathrm{cl}\,W_{1}\setminus S with γ¯2((a2,b2))W1S\overline{\gamma}_{2}((a_{2},b_{2}))\subset W_{1}\setminus S. Let W2W_{2} be the connected open subset of W1W_{1} determined by the Jordan curve γ¯22\overline{\gamma}_{2}\cup\ell_{2}. Arguing exactly as in (C.8), W2SW_{2}\setminus S is open and path-connected, so we can iterate this argument to obtain mutually disjoint embeddings γ¯k:[ak,bk]clW1SΓ1Ω1\overline{\gamma}_{k}:[a_{k},b_{k}]\to\mathrm{cl}\,W_{1}\setminus S\subset\Gamma_{1}\cup\Omega_{1} with γ¯k((ak,bk))Ω1\overline{\gamma}_{k}((a_{k},b_{k}))\subset\Omega_{1} for 1kK111\leq k\leq K_{1}^{1}.

Next, let K11+1\ell_{K_{1}^{1}+1} be another maximal curve with endpoints in Γ1\Gamma_{1}. The same argument as in proving (C.8) implies that Ω1clW1\Omega_{1}\setminus\mathrm{cl}\,W_{1} is path-connected, and so γ(aK11+1)\gamma(a_{K_{1}^{1}+1}), γ(bK11+1)\gamma(b_{K_{1}^{1}+1}) may be connected by a smooth embedding γ¯K11+1:[aK11+1,bK11+1](Γ1Ω1)clW1\overline{\gamma}_{K_{1}^{1}+1}:[a_{K_{1}^{1}+1},b_{K_{1}^{1}+1}]\to(\Gamma_{1}\cup\Omega_{1})\setminus\mathrm{cl}\,W_{1}, that, together with K11+1\ell_{K_{1}^{1}+1}, defines a connected domain WK11+1Ω1W_{K_{1}^{1}+1}\subset\Omega_{1} by the Jordan curve theorem. In addition, WK11+1W1=W_{K_{1}^{1}+1}\cap W_{1}=\varnothing since (2γ¯K11+1)clW1=(\ell_{2}\cup\overline{\gamma}_{K_{1}^{1}+1})\cap\mathrm{cl}\,W_{1}=\varnothing by (C.7) and the definition of γ¯K11+1\overline{\gamma}_{K_{1}^{1}+1}. Repeating the whole iteration procedure for those intervals contained in K11+1\ell_{K_{1}^{1}+1} and then the rest of the maximal arcs, we finally obtain mutually disjoint embeddings γ¯k:[ak,bk]Γ1Ω1\overline{\gamma}_{k}:[a_{k},b_{k}]\to\Gamma_{1}\cup\Omega_{1} with γ¯k((ak,bk))Ω1\overline{\gamma}_{k}((a_{k},b_{k}))\subset\Omega_{1} as desired for 1kK11\leq k\leq K_{1}.

Conclusion of the proof of (C.3): Repeating the Γ1\Gamma_{1} procedure for {Γ2,,ΓIγ}\{\Gamma_{2},\dots,\Gamma_{I_{\gamma}}\} and using the mutual pairwise disjointness of Γi\Gamma_{i}, we obtain mutually disjoint embeddings γ¯k:[ak,bk]clBr(x)S\overline{\gamma}_{k}:[a_{k},b_{k}]\to\mathrm{cl}\,B_{r}(x)\setminus S with γ¯k((ak,bk))Br(x)S\overline{\gamma}_{k}((a_{k},b_{k}))\subset B_{r}(x)\setminus S for 1kK11\leq k\leq K_{1}. We define γ¯:𝕊1Ω\overline{\gamma}:\mathbb{S}^{1}\to\Omega by

γ¯(t)={γ(t)t𝕊1[ak,bk]γ¯k(t)t[ak,bk],   1kK.\overline{\gamma}(t)=\begin{cases}\gamma(t)&t\in\mathbb{S}^{1}\setminus\cup[a_{k},b_{k}]\\ \overline{\gamma}_{k}(t)&t\in[a_{k},b_{k}]\,,\,\,\,1\leq k\leq K\,.\end{cases}

Since γ¯=γ\overline{\gamma}=\gamma outside Br(x)ΩB_{r}(x)\subset\!\subset\Omega, γ¯\overline{\gamma} is homotopic to γ\gamma relative to Ω\Omega. Furthermore, γ¯\overline{\gamma} is piecewise smooth and homotopic to γ\gamma, and so it can be approximated in the C0C^{0} norm by {γj}𝒞\{\gamma_{j}\}\subset\mathcal{C}. However, by (C.4) and the construction of γ¯k\overline{\gamma}_{k}, γ¯S=\overline{\gamma}\cap S=\varnothing, which implies that Sγj=S\cap\gamma_{j}=\varnothing for large jj. This contradicts the fact that SS is 𝒞\mathcal{C}-spanning 𝐖\mathbf{W}, and so (C.3) is true. ∎

References

  • [ACMM01] L. Ambrosio, V. Caselles, S. Masnou, and J.M. Morel. Connected components of sets of finite perimeter and applications to image processing. J. Eur. Math. Soc. (JEMS), 3(1):39–92, 2001.
  • [AFP00] L. Ambrosio, N. Fusco, and D. Pallara. Functions of bounded variation and free discontinuity problems. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, New York, 2000.
  • [All72] W. K. Allard. On the first variation of a varifold. Ann. Math., 95:417–491, 1972.
  • [Alm76] F. J. Jr. Almgren. Existence and regularity almost everywhere of solutions to elliptic variational problems with constraints. Mem. Amer. Math. Soc., 4(165):viii+199 pp, 1976.
  • [BCF13] M. Barchiesi, F. Cagnetti, and N. Fusco. Stability of the Steiner symmetrization of convex sets. J. Eur. Math. Soc. (JEMS), 15(4):1245–1278, 2013.
  • [BM21] J. Bernstein and F. Maggi. Symmetry and rigidity of minimal surfaces with Plateau-like singularities. Arch. Ration. Mech. Anal., 239(2):1177–1210, 2021.
  • [BR97] A. Bhakta and E. Ruckenstein. Decay of standing foams: Drainage, coalescence and collapse. Advances in Colloid and Interface Science, 70(1-3):1–124, 1997.
  • [CCAE+13] I. Cantat, S. Cohen-Addad, F. Elias, F. Graner, R. Höhler, O. Pitois, F. Rouyer, A. Saint-Jalmes, R. Flatman, and S. Cox. Foams: Structure and Dynamics. Oxford University Press, 2013.
  • [CCDPM14] F. Cagnetti, M. Colombo, G. De Philippis, and F. Maggi. Rigidity of equality cases in Steiner’s perimeter inequality. Anal. PDE, 7(7):1535–1593, 2014.
  • [CCDPM17] F. Cagnetti, M. Colombo, G. De Philippis, and F. Maggi. Essential connectedness and the rigidity problem for Gaussian symmetrization. J. Eur. Math. Soc. (JEMS), 19(2):395–439, 2017.
  • [CES22] M. Colombo, N. Edelen, and L. Spolaor. The singular set of minimal surfaces near polyhedral cones. J. Differential Geom., 120(3):411–503, 2022.
  • [CPS20] F. Cagnetti, M. Perugini, and D. Stöger. Rigidity for perimeter inequality under spherical symmetrisation. Calc. Var. Partial Differential Equations, 59(4):Paper No 139, 53, 2020.
  • [Dav14] G. David. Should we solve Plateau’s problem again? In Advances in analysis: the legacy of Elias M. Stein, volume 50 of Princeton Math. Ser., pages 108–145. Princeton Univ. Press, Princeton, NJ, 2014.
  • [DG60] E. De Giorgi. Frontiere orientate di misura minima. Seminario di Matematica della Scuola Normale Superiore di Pisa. Editrice Tecnico Scientifica, Pisa, 1960.
  • [DLDRG19] C. De Lellis, A. De Rosa, and F. Ghiraldin. A direct approach to the anisotropic Plateau problem. Adv. Calc. Var., 12(2):211–223, 2019.
  • [DLGM17a] C. De Lellis, F. Ghiraldin, and F. Maggi. A direct approach to Plateau’s problem. J. Eur. Math. Soc. (JEMS), 19(8):2219–2240, 2017.
  • [DLGM17b] C. De Lellis, F. Ghiraldin, and F. Maggi. A direct approach to Plateau’s problem. J. Eur. Math. Soc. (JEMS), 19(8):2219–2240, 2017.
  • [DM19] M. G. Delgadino and F. Maggi. Alexandrov’s theorem revisited. Anal. PDE, 12(6):1613–1642, 2019.
  • [Dom23] G. Domazakis. Rigidity for the perimeter inequality under Schwarz symmetrisation. 2023.
  • [DPDRG16] G. De Philippis, A. De Rosa, and F. Ghiraldin. A direct approach to Plateau’s problem in any codimension. Adv. Math., 288:59–80, 2016.
  • [DPDRG20] G. De Philippis, Antonio De Rosa, and F. Ghiraldin. Existence results for minimizers of parametric elliptic functionals. J. Geom. Anal., 30(2):1450–1465, 2020.
  • [DR18] A. De Rosa. Minimization of anisotropic energies in classes of rectifiable varifolds. SIAM J. Math. Anal., 50(1):162–181, 2018.
  • [Fed69] H. Federer. Geometric measure theory. Die Grundlehren der mathematischen Wissenschaften, Band 153. Springer-Verlag New York Inc., New York, 1969.
  • [FGS15] M. Focardi, M. S. Gelli, and E. Spadaro. Monotonicity formulas for obstacle problems with Lipschitz coefficients. Calc. Var. Partial Differential Equations, 54(2):1547–1573, 2015.
  • [Fin86] R. Finn. Equilibrium Capillary Surfaces, volume 284 of Die Grundlehren der mathematischen Wissenschaften. Springer-Verlag New York Inc., New York, 1986.
  • [FK18] Y. Fang and S. Kolasiński. Existence of solutions to a general geometric elliptic variational problem. Calc. Var. Partial Differential Equations, 57(3):Art. 91, 71, 2018.
  • [GKJ05] L. Gong, S. Kyriakides, and W.-Y. Jang. Compressive response of open-cell foams. part I: Morphology and elastic properties. International Journal of Solids and Structures, 42(5):1355–1379, 2005.
  • [HP16] J. Harrison and H. Pugh. Existence and soap film regularity of solutions to Plateau’s problem. Adv. Calc. Var., 9(4):357–394, 2016.
  • [HP17] J. Harrison and H. Pugh. General methods of elliptic minimization. Calc. Var. Partial Differential Equations, 56(4):Art. 123, 25, 2017.
  • [JP92] G. Johansson and R. J. Pugh. The influence of particle size and hydrophobicity on the stability of mineralized froths. International Journal of Mineral Processing, 34(1):1–21, 1992.
  • [KHS99] S. A. Koehler, Sascha Hilgenfeldt, and Howard A. Stone. Liquid flow through aqueous foams: The node-dominated foam drainage equation. Physical Review Letters, 82(21):4232–4235, 1999.
  • [KHS00] S. A. Koehler, S. Hilgenfeldt, and H. A. Stone. A generalized view of foam drainage: Experiment and theory. Langmuir, 16(15):6327–6341, 2000.
  • [KMS21] D. King, F. Maggi, and S. Stuvard. Collapsing and the convex hull property in a soap film capillarity model. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 38(6):1929–1941, 2021.
  • [KMS22a] D. King, F. Maggi, and S. Stuvard. Plateau’s problem as a singular limit of capillarity problems. Comm. Pure Appl. Math., 75(3):541–609, 2022.
  • [KMS22b] D. King, F. Maggi, and S. Stuvard. Smoothness of collapsed regions in a capillarity model for soap films. Arch. Ration. Mech. Anal., 243(2):459–500, 2022.
  • [KT17] T. Kagaya and Y. Tonegawa. A fixed contact angle condition for varifolds. Hiroshima Math. J., 47(2):139–153, 2017.
  • [LC99] T. J. Lu and C. Chen. Thermal transport and fire retardance properties of cellular aluminium alloys. Acta Materialia, 47(5):1469–1485, 1999.
  • [LL65] R. A. Leonard and R. Lemlich. A study of interstitial liquid flow in foam. Part I. Theoretical model and application to foam fractionation. AIChE Journal, 11(1):18–25, 1965.
  • [Mag12] F. Maggi. Sets of finite perimeter and geometric variational problems: an introduction to Geometric Measure Theory, volume 135 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2012.
  • [Mil97] J. W. Milnor. Topology from the differentiable viewpoint. Princeton Landmarks in Mathematics. Princeton University Press, Princeton, NJ, 1997. Based on notes by David W. Weaver, Revised reprint of the 1965 original.
  • [MM16] F. Maggi and C. Mihaila. On the shape of capillarity droplets in a container. Calc. Var. Partial Differential Equations, 55(5):Art. 122, 42, 2016.
  • [MNR23a] F. Maggi, M. Novack, and D. Restrepo. A hierarchy of Plateau’s problems and the approximation of Plateau’s laws via the Allen–Cahn equation. 2023. In preparation.
  • [MNR23b] F. Maggi, M. Novack, and D. Restrepo. New convergence theorems for soap films with small volume towards solutions to the Plateau problem. 2023. In preparation.
  • [MSS19] F. Maggi, S. Stuvard, and A. Scardicchio. Soap films with gravity and almost-minimal surfaces. Discrete Contin. Dyn. Syst., 39(12):6877–6912, 2019.
  • [Nov23] M. Novack. On the relaxation of Gauss’s capillarity theory under spanning conditions. 2023.
  • [Per22] M. Perugini. Rigidity of Steiner’s inequality for the anisotropic perimeter. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 23(4):1921–2001, 2022.
  • [Sil05] L. Silvestre. The two membranes problem. Comm. Partial Differential Equations, 30(1-3):245–257, 2005.
  • [Sim83] L. Simon. Lectures on geometric measure theory, volume 3 of Proceedings of the Centre for Mathematical Analysis, Australian National University. Australian National University, Centre for Mathematical Analysis, Canberra, 1983.
  • [SM15] R. Singh and K. K. Mohanty. Synergy between nanoparticles and surfactants in stabilizing foams for oil recovery. Energy and Fuels, 29(2):467–479, 2015.
  • [Tam82] I. Tamanini. Boundaries of Caccioppoli sets with Hölder-continuous normal vector. J. Reine Angew. Math., 334:27–39, 1982.
  • [Tam84] I. Tamanini. Regularity results for almost minimal oriented hypersurfaces in N\mathbb{R}^{N}. Quaderni del Dipartimento di Matematica dell’ Università di Lecce, 1984. Available for download at http://cvgmt.sns.it/paper/1807/.
  • [Tay76] J. E. Taylor. The structure of singularities in soap-bubble-like and soap-film-like minimal surfaces. Ann. of Math. (2), 103(3):489–539, 1976.
  • [WH99] D. Weaire and S. Hutzler. The Physics of Foams. Oxford University Press, 1999.